You are on page 1of 58

The Effects of Wall Roughness on the Particle Velocity

Field in a Fully Developed Channel Flow

by
Michael J. Benson, John K. Eaton

Prepared with support from the


National Aeronautics and Space Administration

Report No. TSD-150

Thermosciences Division
Department of Mechanical Engineering
Stanford University
Stanford, CA 94305-3030

May 2003
ii
Abstract

Previous experimental research has shown that moderate loadings of small, dense particles
can have large impacts in attenuating the turbulence levels of the gas phase in particle-laden
flows. One parameter receiving little attention associated with such flows in fully-developed
channels is that of the wall roughness boundary condition. The objective of this research
was to investigate the nature of the eects of variations in the wall roughness on the particle
velocities.
The experiments were conducted in a well-documented apparatus, rebuilt for the current
work. The facility is the vertical, fully-developed channel flow in air, operated at a Reynolds
number of 13,800 used previously by Kulick et al, Fessler and Eaton and Paris. The final
1.7 m of the development section consisted of two dierent interchangeable sections. One
section, referred to as the smooth wall condition, matches the rest of the development length
and consists of clear acrylic walls. The rough wall section is similarly constructed to the
first, with an overlay of stainless steel wire mesh screen axed to its inner surfaces. The
flow was seeded with monodisperse, spherical 150 m diameter glass particles at a mass
loading of 15%. A single component LDA system was used to obtain flow tracer and particle
streamwise velocity profiles in the wall-normal direction.
The wall roughness condition does not have an impact on gas phase mean velocities in
the tunnel. Gas phase turbulence levels are also quite similar near the wall, while the rough
wall velocity fluctuations exceed those for the smooth wall in the region around the channel
centerplane. Particle velocity PDFs and profiles show a large dependence on the wall
boundary condition. Whereas particles lead the flow in the smooth wall case, they lag the
flow in the rough wall condition, except very near the wall. Throughout the profile, rough
wall particle mean velocities lag their smooth wall counterparts. Particle RMS velocities
are also dierent through the wall-normal profile. Smooth-wall RMS velocities exceed those
of the rough wall nearest the wall, while the opposite is true nearer the channel centerplane.
It was concluded that the rough wall boundary condition enhanced particle-particle and
particle-wall collisions near the wall. It is likely that the wall boundary condition will have
large eects on the gas phase turbulence levels in two-phase flow conditions.

iii
iv
Acknowledgments

The United States Army fully funded the first author through the completion of this work.
The research program was funded by the National Aeronautics and Space Administration
through Grant NAG3-2738.
Grateful acknowledgments go to Dr. Anthony Paris for his previous contributions to
this project, as well as senior PhD candidate Mr. Wontae Hwang, who illuminated many
facets associated with particle-laden flows. In addition, thanks go to Mr. Lakhbir Johal for
his assistance in the rebuilding of several large sections associated with the facility, and Mr.
Ryan Rogers, who provided valuable assistance in data collection.

v
vi
Contents

Abstract iii

Acknowledgments v

List of Figures viii

1 Introduction and Objectives 1


1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Specific Research Background . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 Literature Review on Roughness . . . . . . . . . . . . . . . . . . . . . . . . 3
1.4 Governing Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.5 Specific Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2 Experimental Facility and Techniques 9


2.1 Experimental Facility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1.1 Wind Tunnel Description . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1.2 Particle Description . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.1.3 Flow Tracer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.1.4 Tunnel Control Mechanisms . . . . . . . . . . . . . . . . . . . . . . . 12
2.1.5 Laser Doppler Anemometry . . . . . . . . . . . . . . . . . . . . . . . 12
2.1.6 Data Acquisition System . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2 Experimental Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3 LDA Experimental Uncertainty . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3.1 Single Point Uncertainty . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3.2 Ensemble Statistical Errors . . . . . . . . . . . . . . . . . . . . . . . 16
2.3.3 Velocity Bias . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.4 Flow Qualification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.4.1 Data Comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.4.2 Mean Pressure Measurements . . . . . . . . . . . . . . . . . . . . . . 18

3 Velocity Measurements in the Wall Normal Plane 27


3.1 Single Phase Measurement Results . . . . . . . . . . . . . . . . . . . . . . . 27
3.1.1 Velocity PDFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.1.2 Mean Velocity Profiles . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.1.3 RMS Velocity Profiles . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.2 Two-Phase Measurement Results . . . . . . . . . . . . . . . . . . . . . . . . 29
3.2.1 Velocity PDFs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.2.2 Mean Velocity Profiles . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.2.3 RMS Velocity Profiles . . . . . . . . . . . . . . . . . . . . . . . . . . 30

vii
4 Summary and Conclusions 41
4.1 Summary of Experimental Results . . . . . . . . . . . . . . . . . . . . . . . 41
4.2 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.3 Recommendations for Future Research . . . . . . . . . . . . . . . . . . . . . 43

References 45

viii
List of Figures

2.1 Particle-laden wind tunnel schematic . . . . . . . . . . . . . . . . . . . . . . 20


2.2 Schematic of flowmeter (from Fessler and Eaton, 1995) . . . . . . . . . . . . 21
2.3 Particle-feeder system (from Paris and Eaton, 2001) . . . . . . . . . . . . . 22
2.4 Test section schematic, in cm (from Paris and Eaton, 2001) . . . . . . . . . 22
2.5 Magnified Image of 150 m glass particles . . . . . . . . . . . . . . . . . . . 23
2.6 LDA apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.7 Single phase velocity measurements . . . . . . . . . . . . . . . . . . . . . . . 24
2.8 RMS velocity measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.9 Particle velocity PDF . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.10 Streamwise mean pressure gradient . . . . . . . . . . . . . . . . . . . . . . . 25

3.1 Comparison of gas phase streamwise velocity PDFs . . . . . . . . . . . . . 35


3.2 Gas phase velocity profile comparison . . . . . . . . . . . . . . . . . . . . . 35
3.3 Gas phase rms velocity comparison . . . . . . . . . . . . . . . . . . . . . . . 36
3.4 Comparison of particle phase streamwise velocity PDFs . . . . . . . . . . . 36
3.5 PDFs of streamwise velocities . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.6 PDFs of particle phase velocities near the channel wall . . . . . . . . . . . 37
3.7 Streamwise mean velocity profile . . . . . . . . . . . . . . . . . . . . . . . . 38
3.8 Slip velocity and still-air particle terminal velocity profile . . . . . . . . . . 38
3.9 Streamwise RMS velocity profile . . . . . . . . . . . . . . . . . . . . . . . . 39
3.10 RMS velocity profile comparison . . . . . . . . . . . . . . . . . . . . . . . . 39
3.11 RMS velocity profile crossing point comparison . . . . . . . . . . . . . . . . 40
3.12 Transverse velocity fluctuation intensity for 150 m glass particles at a mass
loading of 0.2 (From Paris) . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

ix
x
Nomenclature
Acronyms

DNS Direct Numerical Simulation


LDA Laser Doppler Anemometry
LES Large Eddy Simulation
PDF Probability Density Function
PIV Particle Image Velocimetry

Roman Symbols

CD coecient of drag
dp particle diameter
h channel half-width
m mass
m mass flow rate
np particle number density
N number of samples
O(n) terms of order n or smaller
P pressure
Reh channel Reynolds number
Rep particle Reynolds number
St particle Stokes number
t Time
U fluid velocity
V particle velocity
x1 streamwise direction
x2 transverse direction
x3 spanwise direction
y+ non-dimensional wall normal distance

Greek Symbols

volume fraction
particle mass loading
k Kolmogorov length scale

xi
fluid viscosity
fluid kinematic viscosity
material density

xii
Chapter 1

Introduction and Objectives

1.1 Introduction
Turbulent, particle-laden shear flows in air occur in a wide range of applications, includ-
ing industrial conveyance processes, pulverized coal combustion, pollution control systems,
fluidized beds, and the environment. As such, they represent an important class of mul-
tiphase flows, requiring detailed understanding to accurately model them. Over the past
several decades, much eort in research has focused on the nature of these complex flows.
To date, no single analytical model confidently predicts all the key properties of these flows.
The complex nature of the flows, and the various parameters influencing them, is not un-
derstood well enough to develop models usable in more general contexts. The very nature
of the interaction between particles and the carrier phase must be determined experimen-
tally for each new flow geometry and parameter set. As such, a great deal of work remains
in this area of turbulence research. To successfully develop models, therefore, two broad
categories of information must be reliably obtained from these two-phase flows. The first
regards the fluid motion. In this category, statistics such as the mean velocity profile and
any modification of turbulence must be captured. In the second, similar statistics regarding
the particle motion in wall-bounded flows is required. Particle mean velocity profile and
spatial concentration data must be collected. Finally, turbulence fluctuation information
must be gathered to fully account for the energy within the flow.

1.2 Specific Research Background


The particular wall-bounded flow studied herein has been studied extensively as part of
the Eaton research group in the Mechanical Engineering Department at Stanford University.
In the past decade, four key experimental studies were conducted in wall bounded turbulent
flows, and their results are summarized here.
In 1989, Rogers and Eaton studied the boundary layer of a turbulent, particle laden
vertical upflow. It was one of the earliest looks at the particle response to turbulence in the
boundary layer and the eect of uniformly dispersed particles on gas-phase turbulence. The
presence of the particles did not change the mean streamwise gas velocity, and they found
reasonably predictable particle mean velocities. However, the particle fluctuating velocities
were substantially greater than predicted. Using a moderate 20% mass loading of copper
particles they found significant attenuation of the gas-phase turbulence.
Kulick and Eaton (1993) prepared one of the more extensive reports on turbulence
modification by particles, using the same fully developed vertical channel flow studied in
the present report. In their study, both fluid and particle statistics were measured through
a range of mass loadings with several types and sizes of fine spherical particles. Like Rogers

1
2

& Eaton (1989), Kulick found no change in the mean streamwise fluid velocity values in the
presence of particles. The fluctuating components of the air velocity decreased as particle
Stokes number, mass loading, and distance from the walls increased. Meanwhile, particle
velocity profiles flattened across the wall normal plane of the channel as the mass loading
increased. Of note, this was the first of the Eaton-group experimenters to notice the strange
eect that particles with larger Stokes numbers lagged the gas-phase in the center of the
channel, while leading the flow near the wall. This is despite the fact that gravity is acting
in the flow direction. Intuition suggests that particles would move more rapidly than the
mean flow.
Fessler and Eaton (1995) used the vertical channel studied by Kulick, modified with a
single-sided expansion. In their study, turbulence attenuation was found in specific regions
of the flow in the same manner as Kulick and Eaton (1993). However, other parts of the
flow, namely the separated shear layer and the redevelopment region behind the step, did
not exhibit any turbulence modification from the unladen state. In addition, they put
considerable eort into the study of preferential concentration eects. They showed that
particles are not uniformly dispersed in the fully developed region of the flow as is often
assumed.
Most recently, Paris and Eaton (2001) further investigated the same vertical channel flow
studied by Kulick (1993) using smaller measurement scales and Particle Image Velocimetry
(PIV) techniques in order to identify more clearly some of the mechanisms at work in the
turbulence dissipation process. Paris (2001) verified the same levels of turbulence attenu-
ation as Kulick (1993). The particle velocities measured by Paris consistently lagged the
streamwise gas phase flow at the channel centerline. In addition, the particle fluctuating
velocity profile in the streamwise direction was relatively flat, similar to data found by
Kulick, and exceeded the fluid phase turbulence levels throughout the profile except nearest
the wall.
Both Kulick (1993) and Paris (2001) found one common, yet surprising result. The
mean velocity of the particles in the streamwise direction at the channel centerplane actu-
ally lagged behind the mean velocity of the flow of the carrier phase. Despite the eects
of gravity, the particles used in those experiments moved slower than expected by an ap-
preciable amount. Numerical simulations, however, do not predict this counter-intuitive
result. Rouson and Eaton (1997) conducted direct numerical simulations on particle ve-
locities. They modeled the vertical channel flow studied by Kulick (1993), using specular
particle-wall interactions at smooth boundaries. The results predicted particle velocities,
using similar particles to those used by Kulick and Paris, that would exceed the mean fluid
flow velocity throughout the channel. Particle fluctuating velocities for copper particles in
the streamwise direction exceed those of the fluid phase near the wall, while these fluc-
tuating velocities are nearly equal in the channel centerline. This eect, which shows a
general decrease of fluctuating velocity as one moves from the wall to the channel center, is
a marked contrast from the flat profiles measured by Kulick and Paris.
As an attempt to explain this particle behavior, Paris (2001) suggested four potential
modes whereby particle mean velocities might be reduced. They include insucient chan-
nel development length, particle migration, particle-wall collisions, and particle-particle
Chapter 1. Introduction and Objectives 3

collisions. Of these four, they discount the first, due to the large particle residence time in
the upstream development section versus the particle time constant. The remaining poten-
tial contributors are all closely linked in terms of the physical processes by which they occur.
Increased transverse particle velocities due to particle-particle collisions were deemed the
most likely cause of reduced particle streamwise velocity because the particle concentration
seems to play a large role in the amount by which a particle lags the mean flow. Including
the influence of surface roughness eects in the overall flow analysis was one mechanism
suggested by Paris and Eaton to help explain some of the more surprising aspects of the
particle behavior.

1.3 Literature Review on Roughness


The eects of roughness in single phase flows have been studied in some detail for a
number of types of flow geometries. For example, roughness types were described in Perry
et al (1969), classifying k and d type roughness, each with diering eects in regions close
to a wall or surface. In Perry et al (1987) it was concluded that the eects of roughness
were small at large distances from the surface compared with the size of roughness elements.
Raupach et al (1991) summarizes a review of several related studies with similar conclusions.
However, in Krogstad et al (1992) significant eects of wall roughness were seen well outside
this wall region. Further, Smalley et al (2002) identified problems with the generalized use
of the roughness functions used in so many of these and other similar studies. Thus,
roughness in single phase flows continues to be a subject of contention despite numerous
previous studies.
On the other hand, roughness eects in two-phase flows have been studied far less.
With the discovery of the attenuation of turbulence in specific two-phase flow regimes in
the past few decades, little has been concluded with respect to the eects of wall roughness,
especially regarding specific particle transport phenomena. However, broad categories have
been introduced by which these eects may be investigated. Four categories of important
eects include particle concentrations, particle interaction with the carrier phase, particle
interactions with walls, and particles interacting with other particles.
In an experimental study of a downward directed turbulent pipe flow, Varaskin et al
(2000), used 50 m spherical glass particles in an air flow through a 46mm diameter by
2.5m long stainless steel pipe. They ranged their experiments through three regimes of
mass loading, from very dilute, to moderately dense loadings. In each case, they measured
axial mean velocities of the particulate phase that exceeded the corresponding gas phase
mean velocities throughout the profile. In terms of fluctuating velocities, those measured
for the particles increased as the wall was approached. This trend showed increasingly sharp
gradients near the wall. As the particle concentrations increased, so did this fluctuating
velocity. In general, increasing particle concentration reduced overall particle streamwise
velocities, and increased velocity fluctuations in near wall regions.
In another study, Kaftori, Hetstroni, and Banerjee (1998) used near neutral-density par-
ticles in a horizontal water channel. They studied particle mixing and concentration eects
4

on their turbulent flow, which was in the regime where carrier turbulence is attenuated by
particles. Of note in their experiment, they identified a correlation between the particle
concentrations with the behavior of flow over a rough wall. They were careful to identify
this eect only in terms of a non dimensional sand roughness. The overall eect of particle
concentrations in the flow act in such a way as to produce eects similar to that of a rough-
wall. However, the overall concentration eect, in terms of this equivalent wall roughness
parameter, remained quite small, and was still within the hydraulically smooth regime for
their experiments. Using particles ranging from 100 - 900 m, they found that larger parti-
cles increased this eect. Despite the resemblance of the two eects, they were unsure of the
similarity of the physical mechanisms causing the eect. Finally, they identified streakline
concentration behavior of the particles. Highest particle concentrations were found in low
velocity regions, such as near the walls, with concentration values exceeding those of other
regions by one to two orders of magnitude.
Recently, Tanaka, Maeda, and Hagiwara (2002) conducted numerical studies in a ho-
mogeneous turbulent shear flow. They performed their study using a particle loading in
the very dilute range, where inter-particle collisions had no eects, modeling their flow
with the relatively traditional two-way coupling analysis. They found that gravity plays
an important role in the particle interaction with the carrier phase. In weak gravity condi-
tions, particles actually increased turbulence, but in strong gravity, their overall eect was
to attenuate turbulence. They ascribed the dierence to the location of particle clusters
within the flow. In the weaker gravity conditions, the particle clusters were located in such
a way as to increase the downward flow and enhance Reynolds shear stress. However in the
stronger gravity regimes, particle clusters shifted to locations where drag dissipation eects
attenuated the flow turbulence. This analysis blended the importance of the particulate
interaction with the carrier phase with the eects of local concentrations of particles.
Important studies have also focused on how particles interact with the walls in turbulent
flows. In an experimental and numerical study conducted by Frank et al, (1993), a square
horizontal channel was used to shoot particles at walls of dierent consistencies. They
studied the resulting coecients of restitution and the dynamic friction coecients in an
attempt to better predict what happens to particles after contact with walls. Generally, they
found that particles in the presence of rough walls rebound at a fairly consistent mean value
of 65 degrees from parallel. Using these values, their numerical results agreed reasonably
well with their experiments. In each of their comparative profiles, the particles did lag the
mean flow by a slight percentage. They attributed this lag to three-dimensional flow eects
developed in their square cross-section.
Sommerfeld and Huber (1999) did a similar study of the particle-wall interaction eects.
Glass and quartz beads with diameters between 100 and 500 m were used. As in Frank et
al, they focused on particle rebound properties. Wall roughness eects were seen to depend
on particle size and geometrical eects of the roughness elements. In general, particles
were seen to rebound at increasingly large angles as the relative roughness of the surface
increased. Non-sphericity was also a factor in the rebound characteristics of the particles.
Using the coecients of restitution and friction from their experiments, they compared their
resulting rebound angles from a variety of impact angles for polished steel and rubber walls,
Chapter 1. Introduction and Objectives 5

with excellent agreement.


Yamamoto et al, (2001) conducted a large eddy simulation where they investigated
inter-particle collisions. Throughout much of the early literature regarding two-phase flows,
inter-particle collisions are disregarded for reasonably disperse mass loadings of particles.
In this study, the conditions of the experiment of Kulick and Eaton (1993) , were modeled
to see the impact of considering these inter-particle eects at mass loadings of 20%. In
terms of comparison of particle mean velocity profiles, they found that including the eects
of inter-particle collisions yielded reasonably good information regarding velocity gradients.
However, they were less successful in matching the magnitudes of the velocity, which were
generally lower than those calculated with the collisions taken into account. They attribute
this dierence to the particle wall interaction on account of the wall material utilized in the
experiment. As regards particle fluctuating velocities, they found that in the streamwise
direction particle-particle collisions could be disregarded. However, in the wall-normal
component of fluctuating velocities for copper particles, inter-particle collisions were seen
to have a large eect and raised the fluctuating velocity levels significantly.
Fohanno and Oesterle (2000), also investigated particle-particle collisions, using both
experimental and numerical methods. They used relatively large 3mm glass particles in
a vertically convergent flow to force particle-particle collisions at desired measurement lo-
cations. Due to the unique experimental setup, the particle-particle interaction increased
as the walls of the test facility converged. Similar to Yamamoto et al. (2001), Varaskin
et al. (2000), Kaftori et al. (1998), and others, they again found the highest particle con-
centrations in the low velocity regions near the wall. The initially high values of particle
fluctuating velocities in the transverse direction were attenuated as the particle collisions
increased, while those same fluctuating velocities in the streamwise direction remained rela-
tively unchanged. This served to increase the overall homogeneity of the particle fluctuating
velocity in the channel. The net result is that the inter-particle collisions counter the eects
of preferential concentration and reduce fluctuating velocities across the channel, resulting
in a particle flow which, tends towards a simple vertical flow.
Finally, in a detailed study that addresses particle concentrations, as well as particle-
gas, particle-wall, and particle-particle interactions, Kussin and Sommerfeld (2002) used
a horizontal channel with exchangeable stainless steel wall plates. They tracked particle
velocity characteristics and particle concentration profiles in regimes with walls constructed
of both low and high roughness materials. In addition, they varied conveyance velocity, mass
loading, and used particles ranging from 60-190 m in size. They found in the streamwise
direction that increasing the wall roughness decreased the particle mean velocities across
the channel, increased the particle fluctuating velocities, increased the level of isotropy of
particle concentrations (for the 100 m particles at least), and increased the already well
established eects of mass loading. These changes were primarily attributed to the wall
roughness increasing the frequency of inter-particle and particle-wall collisions.
6

1.4 Governing Equations


When considering the particle phase, it is important to begin by establishing some of the
primary equations to be used in describing flow properties. One such property important
to define, then, is the particle Reynolds number. This non-dimensional property represents
the ratio of the inertial to the viscous forces in the vicinity of the surface of the particle.
f | V U | dp
Rep = (1.1)

where, V U represents the slip velocity between the particle phase and the fluid phase. For
very small values of Rep , the Stokes regime conditions are met. For Rep values larger than
unity, but under approximately 200, the empirical approximation of Shiller and Naumann
(1933) yields a coecient of drag that matches well with experimental data:
24
CD = [1 + 0.15Re0.687
p ] (1.2)
Rep
The full particle equation of motion for Stokesian particles in wall bounded flows includes
several terms. These terms account for gravity forces, viscous drag, a Stokes force, a force
due to pressure gradients in the surrounding fluid, an added mass force described by
(Kulick 1993) as the force required to accelerate the fluid displaced by a spherical particle,
and the Basset history force. However, in the case where the particle density is much greater
than the fluid density, as the gas-solid flows in the present experiment are, terms of order
f /p may be neglected. In this case, the particle equation of motion becomes

dVi 1
= gi (Vi Ui ) (1.3)
dt p
In the case of non-Stokesian particles, particle inertial eects must be accounted for,
and the resulting dierential particle equation of motion then becomes (using the empirical
formula for the CD from above):

dVi 18
= gi [1 + 0.15Re0.687
p ](Vi Ui ) (1.4)
dt p d2p
This equation can be integrated forward in time until the terminal slip velocity (Vi Ui )
is reached. This terminal velocity, which neglects eects associated with any cross stream
mixing of the particles, represents the fastest the particle should travel for the given ex-
perimental configuration. For example, given a constant fluid velocity of 10.5 m/s, the
non-Stokesian terminal particle velocity for a 150m sperhical glass particle and the above
assumptions would be 11.48 m/s, reached in approximately 0.5 seconds.

1.5 Specific Objectives


The objectives of this report are straightforward. First, a well documented fully devel-
oped vertical channel apparatus is reconstructed for use. Walls made of aging particleboard
Chapter 1. Introduction and Objectives 7

are replaced with acrylic Plexiglas, optical access is improved, and components checked to
ensure everything operates in such a way as to make the entire experiment reproducible.
This included developing an LDA system capable of accurately measuring particle velocity
statistics. In addition, the final two meters of the development length are built with three
exchangeable wall plates including smooth walls made of plain acryclic, medium-rough walls
made of particleboard, and rough walls consisting of wire mesh screen attached to acrylic.
The second objective of this work is to measure particle mean velocities and streamwise
turbulence intensity of each of these three roughness regimes and compare and contrast
their statistics. It is estimated that this second objective will definitively answer whether
wall roughness is the primary reason by which mean particle velocities are reduced.
8
Chapter 2

Experimental Facility and Techniques

This chapter discusses the experimental apparatus and the data measurement techniques
used in this study. Some qualifications of the techniques are presented, and estimates and
analysis of the sources of error are discussed. Many of the details contained in this section
are similar to works done by Paris (2001), Fessler (1995), and Kulick (1993), and thus
details are provided here only when there are significant dierences relative to the previous
works.

2.1 Experimental Facility


The experimental facility consists of an open-loop, vertical, high aspect ratio, fully
developed channel wind tunnel, a particle feeding apparatus, a Laser Doppler Anemometer,
and associated data acquisition and experimental control devices.

2.1.1 Wind Tunnel Description


Figure 2.1 shows a general schematic of the vertical wind tunnel facility, and illustrates
some of its major aspects. It is a very tall facility, spanning from the building roof through
each of its three floors. A large upper platform, mounted 7.0 m from the base of the facility
and of entirely new construction, is secured in multiple areas to the structural supports of the
building. The tunnel sections themselves are secured throughout the vertical span, using
rubber vibration-damping pads, to intermediate supports, and are completely immobile
during experimental runs at even the highest speeds.
Air enters the wind tunnel through a 0.3 m3 plenum attached to the upper platform.
The air passes through a 0.25 m pleated air filter which is changed regularly. A flow meter
connects directly to the exit of the plenum. This flow metering system is the same custom
venturi-type meter described by Fessler (1995). It is constructed of Renshape 450 rapid
prototyping material, with dimensions as shown in figure 2.2. A temperature compensated
pressure transducer (Setra model 239) samples the pressure dierence between the inlet
plenum and a set of three parallel pressure taps 30.5 cm into the length of the flow meter.
The pressure transducer in the flow metering section provides the data whereby the mass
flow rate of the air is measured and controlled during tunnel operation.
The flow metering section connects with a flexible rubber boot to a New York Blower,
size 124 compact G.I. CCWBH fan, which draws the air through the filter and flowme-
ter and forces it through the tunnel. The blower is driven by a 5-hp AC motor with
eddy-current clutch, and controlled with an Eaton Dynamatic Model 4000 speed controller
system. The blower/motor system is supported on four vibration-damping feet, and secured

9
10

to the wooden platform. In this configuration, the blower can create tunnel velocities in
excess of 15 m/s.
The air exits the fan through a rectangular diuser and enters a 25.4 cm diameter sheet
metal duct. It travels 2.5 m through the duct, and enters a particleboard duct at the top of
the wind tunnel. Over a horizontal distance of 70 cm, this section turns the flow 90 degrees
from a 40 cm x 25.6 cm rectangular section, into a 40 cm x 45.7 cm rectangular cross
section, 0.87 m long. The header-type 90 degree bend replaces a flexible pipe connection
on the original apparatus. Flow-conditioning grids are placed in four locations throughout
this section. A clear acrylic 40 cm x 45.7 cm rectangular section, 0.82 m long, follows
the particleboard ductwork. Particles are introduced into the flow in this region. A flow-
conditioning grid follows the particle loading system. A wire mesh screen follows this grid,
and a second flow-conditioning grid follows the screen. The grids and screen act to uniformly
disperse the particles in the flow.
The particle loading system, as shown in figure 2.3, is identical to that used by Paris
(2001). It consists of a large number of drainage buckets connected to two sets of conveyer
belts that rotate across the breadth of this loading section. The two sets of drainage buckets
are filled through a particle hopper mechanism. Pressurized air starts the flow of particles
through drainage tubes into the rapidly moving buckets. The buckets then drain steadily
across the tunnel to the opposite side where a matching feeding mechanism refills them
for the return trip. The buckets are constructed with inlets that are much larger than the
outlets, allowing them to be filled rapidly while draining more slowly. The particle hoppers
each hold 17 L of particles, so a typical experimental run can last for nearly three minutes
with a constant mass flow rate. The belts are moved by a 1/12 hp Boston Gear DC motor
with an RB1B Beta II speed controller. Adjusting the speed of the controller can increase
or decrease the resultant mass flow rate of particles in the channel.
The two-phase flow next enters a 10:1 contraction that smoothly transforms the 40
cm x 45.7 cm cross section into the high aspect ratio channel cross section with interior
dimensions of 4.0 cm x 45.7 cm. A 5 cm long aluminum Hexcel honeycomb section is pressed
into the end of the contraction to minimize the cross stream velocity of the particles exiting
the contraction. After exiting the honeycomb, three newly constructed 1.27 cm thick acrylic
channel development sections follow. Each section is seamlessly connected with the previous
through a flange and gasket connection, and the total height is 5.2 m (260 channel half
heights), with constant interior dimensions of 4.0 cm by 45.7 cm. Support ribs throughout
the development sections assure that the overpressure in the tunnel does not significantly
bulge the channel walls. The transparent nature of the acrylic sections allows for one
to visually see the particles from the exit of the honeycomb cells at the contraction exit
through the test section. Static pressure taps are emplaced every 30.5 cm throughout the
development section.
The final of the three 1.7 m long development sections is an exchangeable section that
is relatively easy to replace. There were three versions of the replaceable section that could
be exchanged to investigate the eects of dierent wall boundary conditions. The first
version was made of smooth acrylic. It is smooth both to the touch and optically, with no
measureable roughness. The second version was the rough-wall development section. This
Chapter 2. Experimental Facility and Techniques 11

was another acrylic section, but square-mesh wire screen was bonded to both the 45.7 cm
wide walls to make a surface with large regular roughness. The screen was manufactured
from 0.25 mm diameter wire at a grid spacing of 1 mm. The screen was bonded to the
acrylic using a 3M 90 high strength spray adhesive. The third version had one of the 45.7
cm walls covered with the same screen, and the opposite was manufactured with 12.7 mm
thick particleboard.
Particle velocities in the development section were approximately 10 m/s near the chan-
nel centerline, and slower elsewhere. Therefore, even a fast moving particle took approxi-
mately 170 ms to traverse the length of the final development section. Most particles took
longer. Also, the length of the final development section corresponds to 85 channel half-
widths. Therefore, it was believed that the final development section was suciently long
for the flow to become fully developed prior to entering the test section.
Figure 2.4 shows the dimensions and layout of the newly constructed test section. It has
a total length of 63.5 cm, and is optically transparent. Regardless of the wall conditions
in the final development section, the test section retained its smooth acrylic walls required
for optical access. Based on the average particle speeds in this region, the particles have an
average residence time of less than 30 ms before reaching the measurement volume. The test
section has five static pressure ports, each 2.54 cm apart, as well as a column of instrument
ports 5 cm from the channel centerline for devices such as pitot probes and temperature
sensors. The test section has optically clear glass walls in the middle 20 cm of each side of
the 4 cm wide side walls. These walls are clearer than the acrylic, so oer improved optical
imaging and laser transmittance. In addition, a 20 cm square glass window in the center of
one 45.7 cm channel wall facilitates the performance of the collection optics. The opposite
side test section wall has a matching sized, removable acrylic window to allow for access
into the interior of the channel for purposes of laser alignment and any required cleaning.
Of note, the entire test section support frame was emplaced using a laser level device. The
entire tunnel was then secured to the frame ensuring that it was square with the fixed frame,
thereby validating the vertical orientation of the entire apparatus.
At the exit of the test section, a custom designed sheet metal contraction smoothly
transitions the rectangular duct flow into a 10 cm diameter PVC pipe, which turns through
a 90 degree elbow, travels horizontally for 1.4 m, goes through another 90 degree elbow,
travels vertically for 2.4 m, bends a third time, and enters an Environmental Industrial
Products model FK912D1 NF12 cyclone separator. The cyclone separator extracts the
particulate phase from the gas phase, exhausting the air to an exit 7 m above and 10 m
horizontally through the building roof, while dropping the particles into a stainless steel
tote directly beneath the cyclone. At the end of each experimental run, the mass inside
the tote was weighed. Dividing the mass by the duration of the particle flow gave the
particulate mass flow rate. The particles were reused several times, and were brought back
to the upstairs loading platform using a 9 m electric chain hoist.
12

2.1.2 Particle Description


Glass beads of the same manufacture as those used by Paris (2001) were utilized through-
out this experiment. The glass particles were nominally monodisperse spherical beads with
diameters 150 m. Particle sizing was done by Vortec Products of Los Angeles, CA, with
a standard deviation of 10 m. A magnified photograph of the particles, highlighting their
uniform size and sphericity is shown in figure 2.5. Table 2.1, adopted from Paris (2001),
summarizes the particle properties and some statistical quantities associated with the par-
ticles in the smooth wall configuration.

2.1.3 Flow Tracer


Flow tracer particles were required for the gas-phase measurements used to qualify the
apparatus. The flow tracers were created using a ROSCO Delta 3000 fog machine, which
vaporizes fog oil, that subsequently condenses into nominally 0.25 m diameter polydisperse
particles. The density of the flow tracers could be adjusted using a simple connection to
the fog machine and regulating the input voltage. For ease and accuracy of mass flow
calculations, the fog machine was placed inside the plenum chamber of the air intake system.

2.1.4 Tunnel Control Mechanisms


The tunnel operation was automated as much as possible to provide consistent experi-
mental conditions for the numerous runs. There were two primary control mechanisms that
accomplished these tasks. The first was the calibration of the upstairs pressure transducer.
Using a pitot tube - manometer system at the test section and a Labview program, the
single phase centerline velocity of the tunnel was adjusted through a wide range of condi-
tions. The resulting voltage readings from the upstairs pressure transducer, gathered from
an analog output line and read by a data acquisition board on a Pentium 3 computer, were
plotted against the manometer readings. The result was a linear plot where the slope and
intercept could be used in a dierent Labview program to control the tunnel speed. The
tunnel speed automatically corrected itself to be accurate within 0.01 m/s of the desired
tunnel centerline velocity. As particles were introduced into the flow, and the net flow speed
reduced, the tunnel would take approximately 45 seconds to correct itself to the desired flow
speed before any measurements were taken.
The speed controller system of the tunnel depended on the output from the Labview
program sending voltages to the Eaton Dynamatic 4000 speed controller connected to the
blower. The resistance and capacitance potentiometers inside the speed controller were
adjusted to allow the 10 Volt range of the Labview program access to appropriate tunnel
speeds throughout a wide range of mass loadings.

2.1.5 Laser Doppler Anemometry


A dual beam, single component, Laser Doppler Anemometer operating in side scatter
mode was set up for use in the experiment. Much of the system was built from a range of
Chapter 2. Experimental Facility and Techniques 13

TSI Inc. products used in Fessler (1995), and Kulick (1993). Figure 2.6 shows the basic
components of the LDA system, and Table 2.3 describes the primary components for the
complete system. The transmitting optics assembly was mounted on a flat rail system, which
in turn was controlled by a traverse. The traverse, oriented parallel to the Y component of
the tunnel (that is, in the wall-normal direction), was powered by a 1/6 hp APM916AT-1
Boston Gear DC motor, controlled by a Boston RB1R Beta II speed controller.
The elliptical measurement volume was estimated, based on fringe spacing and the
number of fringes in an oscilloscope trace of an average signal, at 75 x 500 m, with an
average depth estimated at 75 m. This fairly small measurement volume, compared to the
150 m glass particles, yielded excellent data rates during all experimental runs and for all
of the mass loadings tested in the experiment.
The laser used in the study was a Lexel Model 95 Argon-Ion laser. Initially, the laser
was operated in its standard multiline operational mode, where green and blue wavelengths
dominated. To increase the data aquisition rate the laser was placed into single line mode
by adding an etalon device into the laser head. The 488 nm wavelength (blue) used was
operated at a continuous power of 1 Watt.
Beams were oriented in the X Z plane of the tunnel, entering through the narrow channel
wall, and crossing near the center of the 45.7 cm wide channel dimension. This allowed for
the long axis of the measurement volume to be oriented in such a way as to provide the best
spatial resolution in the wall normal plane of the tunnel. Only streamwise components of
velocities were measured with this apparatus, and all profiles subsequently referred to are
in the wall-normal direction.
Once data were taken at one location, the transmitting optics were moved by the tra-
verse to the next profile location. There were 8 meaurement locations within the channel,
including centerline (20 mm from wall), 10 mm, 8 mm, 6 mm, 4 mm, 3 mm, 2 mm, and
finally 1 mm from the wall. After moving the traverse so that the beam crossing point
coincided with the appropriate profile location, the collection optics was then moved on a
Melles Griot optical platform until they were again refocused on the measurement volume.
For all experiments, the collection optics were focused such that the maximum data rate
was achieved when using a TSI model 9302 atomizer oriented into the measurement volume.

2.1.6 Data Acquisition System


The collection optics for the system consisted of the 9055X0341 Dantec Corporation 80
mm focal length optical collector. This device was specially designed to directly attach to
a TSI model 9162 photomultiplier tube. The photomultiplier was supplied by a TSI model
9165 1/4 ampere power supply. The output signal from the photomultiplier was sent to
a TSI IFA655 digital burst correlator, which extracted the frequency, phase, burst transit
time, and burst arrival time information from the raw electrical signal. A bandpass filter
removed the Doppler pedestal, amplified and clipped the signal, and conducted high and low
pass signal filtering for the desired velocity range. A 1-10 MHz range for the signal filtering
was used, which for the calculated fringe spacing of 1.78 m retained measured velocities
in the 1.78-17.8 m/s range. The accuracy of the digital burst correlator was cross-checked
14

by saving several hundred measurements in a digital storage oscilloscope, and calculating


the resultant velocities by hand. The velocity comparison made it apparent that they were
identically accurate processes.
The output from the digital burst correlator was sent to an InfoGold P100 computer
with installed TSI Inc., FIND software through a direct memory access cable. The data
stored arrival time information, as well as calculated velocities based on the fringe spacing
parameter inputs. The data were then written to text files for analysis in other programs.
Of note, at no time were both flow tracers and particles in the tunnel at the same time.
This is a marked departure from the earlier studies. Since this study focuses on the behavior
of the particles themselves, the flow tracer data merely help to qualify the techniques, as well
as provide a comparison for the behavior of the particles. One issue addressed by Kulick,
Fessler, and Paris was that of cross-talk, which involved flow tracers being identified as
particles. This cross-talk phenomenon was not a factor in these experiments.
All data points in this study represent a minimum of 4000 individual particle velocity
measurements. For all profiles contained in this report, a minimum of four experimental
runs was conducted, and the statistically stationary results averaged. In the case of flow
tracer data, points actually represent ensemble averages of four runs, each with 20000 flow
tracer velocity measurements.
The most basic flow qualification was to take a computer cooling fan, with a manu-
facturer specified rotational rate, and attach a capillary tube to the fan blades. As the
device rotated, the capillary tube was positioned in such a way as to cross through the
measurement volume of the LDA. Both the digital Burst Correlator and the Digital Storage
Oscilloscope were used to compute the velocities of the capillary, which were expected to be
both constant, and could be calculated analytically using the rotational speed of the fan.
Identical, consistent velocities were achieved when comparing the analytical velocity with
both the oscilloscope and the IFA 655.

2.2 Experimental Procedure


The general procedure for an experimental run is detailed below. The wind tunnel is
turned on, and the lab computer immediately assumes control of its speed. The particle
feeder belts are maintained at an appropriate belt speed to produce the desired mass loading
for the experiment. Despite being unclamped, drainage tubes from the particle hoppers do
not drain immediately into the belts of the particle-feeder, as the air pressure in the channel
is approximately 1.3 kPa above the ambient pressure. To begin the flow of particles into
the tunnel, compressed air is supplied to both particle hoppers. A stopwatch is started
when the compressed air line and buckets are turned on. When the air-line is activated, the
particles begin to fill the buckets, and particle loading begins.
At this time, the lab computer compensates for the increased flow resistance and the
tunnel rapidly converges on the desired flow speed. Normally, this process takes about
30-45 seconds depending on the tunnel velocity. Once the flow becomes steady, the data
acquisition computer starts recording data for 120 seconds, or until 10,000 records are
Chapter 2. Experimental Facility and Techniques 15

stored. The recording of data can also cease at the discretion of the experimenter. For
most two-phase experiments, approximately 5000 velocity measurements are aquired.
At the conclusion of the experimental run, the particle-feeder belts are switched o and
the stopwatch is stopped. The total time interval for the particle loading can be accurately
recorded. A Pelouze 4010 heavy duty shipping scale with a resolution of 0.05 kg is used
to weigh the particles collected in the steel tote. The final step in the experimental run is
to shut o the compressed air supply to the particle hoppers. The total elapsed time for an
experimental run is approximately 120 seconds. The size of the particle hoppers limit the
flow of particles to a maximum of 150 seconds. Over the course of this study, laboratory
temperatures varied by less than 5 degrees Celsius and atmospheric pressure readings varied
by less than 0.3 kPa.

2.3 LDA Experimental Uncertainty


This section of the report covers the primary sources of uncertainty aecting the data.
It specifically addresses those issues inherent to LDA measurement systems.

2.3.1 Single Point Uncertainty


One such issue, of only nominal importance because of its use in the analysis, is how
well the flow tracers actually follow the flow of the air. Each of the three previous studies
on this facility, conducted by Kulick, Fessler, and Paris, used flow tracers on the order of
1 micron or below. As this experiment also used flow tracers within that range, the same
conclusions can be drawn; namely, error associated with flow tracers not following the flow
are virtually negligible.
Secondly, the actual alignment of the LDA apparatus itself could present a source of
error. As the optics themselves are calibrated by the manufacturer, errors associated with
these are also considered negligible. Critical elements, such as beam spacing and beam
crossing angles, were individually measured to confirm their calibrations. At most, the
errors in the beam crossing angle could be estimated at 0.1 degrees, which would provide
a constant scaling error for all measurements, and thus have a very limited eect on the
trends shown by the data.
Next, errors associated with near wall measurements could be influenced due to mis-
alignment of the measurement volume. This would be most dramatic in the near wall region,
where the largest velocity gradients are found and the influence of the spanwise velocity
component could contaminate the streamwise velocities. However, measurements in this
report did not approach closer to the wall than 1 mm, so the near wall eects due to this
phenomenon are limited.
Finally, in terms of the potential of having multiple particles within the measurement
volume, it might be conceivable that their Doppler signals would sum to produce a signal
whose frequency could not be measured accurately. However, as the size of the particle is
nearly 30% of the total measurement volume size at its elliptical center, the likelihood of this
error drops substantially. Combining this with the Gaussian distribution of the intensity of
16

the beams further supports this conclusion. Finally, the digital burst correlator thresholds
all Doppler burst records that are substantially dierent than their neighbors, a feature
which all but eliminates chances of this error.

2.3.2 Ensemble Statistical Errors


When analyzing useful statistics from the LDA data, the individual velocity measure-
ments were ensemble averaged over a number of experimental realizations. Since the number
of data points is finite, uncertainty is introduced to the statistical quantities. The contri-
bution of statistical uncertainty to the velocity mean and variance may be estimated from
the Student-t distribution (Bendat and Piersol, 1971).
The uncertainty in the mean for a given confidence level is given by

st st
x x < x + (2.1)
N N
where x is the true mean, x is the sample mean, N is the sample size, s is the sample
standard deviation, and t is the area under the Student-t distribution for the given value of
N (t=1.96 for 95% confidence as N ). The uncertainty in the variance calculation is

(N 1)s2 2 (N 1)s2
x < (2.2)
2(N1); 2(N1);1
2 2

where x2 is the true variance and 2N; is the chi-squared distribution. For large values of
N, the chi-squared distribution is approximated by
" s #3
2 2
2N; = (N 1) 1 + z (2.3)
9(N 1) 9(N 1)

where z is a function of the confidence level for the chi-squared distribution (z =1.96 for
95% confidence as N ).
LDA velocity measurements used a minimum of 4000 samples. For a 95% confidence
interval for the particulate phase, this uncertainty varies from 0.6% at the centerline to 1.1%
near the wall. In the case of gas phase statistics, where 20,000 samples were taken, this
corresponds to uncertainties varying from 0.2% to 0.5% for the same confidence interval.
The corresponding uncertainty using the chi-squared distribution for the rms fluctuating
velocities in the particle phase was on the order of 2.5% at the channel centerline, and
reduces as the wall is approached.
The error associated with the digital burst correlator is extremely low, as the manufac-
turer provides a sample resolution accuracy on the order of 0.05%.
Perhaps a more likely source of error to consider is the actual positioning of the beams
within the channel. The beams were initialized at the channel centerline, which corresponds
to 20 mm from each wall. This distance was measured with a ruler with 0.5 mm accuracy.
Each subsequent wall normal distance was measured with this same technique and same
Chapter 2. Experimental Facility and Techniques 17

ruler. Each data point in the profile represents an average of at least four experimental runs,
which tends to reduce this type of location uncertainty. Since one individual measured all
wall normal distances of the beams for each experiment, the error in this measurement is
estimated as very low, on the order of 5% as a conservative estimate. As this error represents
uncertainty in the actual location within the profile for a given data point, each point has
an equal likelihood of actually representing a point less than one particle diameter on either
side of its stated location, which incurs very minimal eects on the analyses in this report.

2.3.3 Velocity Bias


Because, for a given time interval, high velocity particles are more likely to traverse the
measurement volume than low velocity particles, there may be a tendency toward a velocity
bias in LDA mean velocity measurements. However, it is important to note that this bias
does not imply that measured velocities have an associated error, rather it implies only that
higher velocity particles are more frequently sampled than lower velocity ones. In Kulick et
al (1993), where the same experimental conditions and apparatus were used, this bias was
estimated at 1%, and no correction was applied to the given data. Furthermore, in the case
of glass particle measurements, the desired quantity is the actual particle average velocity,
so there is no bias.

2.4 Flow Qualification


The flow qualification used in this experiment was reasonably extensive. This is due
to the fact that a great deal of data already have been aquired with the apparatus by the
previous researchers referred to throughout this report.

2.4.1 Data Comparison


While the comparison between an analytical solution and experimental results is nom-
inally the best flow qualification technique, another good comparison for this experiment
can be expected from plotting the single-phase statistics of previous workers compared to
current data. However, since the tunnel changed significantly since the results were taken,
the single-phase data from both smooth walls and rough walls is plotted with those of Kulick
and Eaton (1993) and Paris and Eaton (2001). The four data sets comprising this unladen
comparison are shown in figure 2.7. Second, the corresponding unladen, centerline, RMS
velocity fluctuations are shown, comparing the rough-wall case and the data of Paris (2001)
for mass loadings of approximately 10%, in Figure 2.8. Finally, a rough-wall comparison
with Paris (2001)of the PDFs of the centerline particle velocities with similar mass loadings
is shown in Figure 2.9.
Table 2.2, adopted from Paris (2001) summarizes some relevant flow properties associ-
ated with the smooth wall configuration. Of note, the measured wall shear velocity in this
table is somewhat less than that used by the previous experimenters, as might be expected
when changing the composition of the entire development section from particleboard to
18

acrylic. This change in turn aects the values of dependent flow parameters in the fluid
flow parameter table.

2.4.2 Mean Pressure Measurements


The pressure taps that were placed throughout the development section of the channel
and into the channel test section were used to measure the streamwise variation of the
wall pressure along the length of the channel. With the tunnel running at its normal
experimental velocity, a calibrated Setra Model 239 pressure transducer (250P a range)
was used to sample 10,000 data points at each of several locations along the height of
the channel. A Labview program aquired the data, and plotted the normalized results.
Figure 2.10 shows the results for single phase channel flow at a channel Reynolds number
of approximately 13,800.
The wall shear stress was calculated by

dP
w = h (2.4)
dx
and the shear velocity is
r
w
U = (2.5)

A least squares fit of the data yields a shear velocity of 0.47 m/s, in fair agreement with
the value of 0.49 m/s measured by Kulick and Eaton (1993). Paris and Eaton (2001) showed
an unladen wall pressure gradient of 12.2 m/s2 , as compared to the pressure gradient of
12.1 m/s2 measured in this report.
Chapter 2. Experimental Facility and Techniques 19

Table 2.1. Particle parameters

density (kg/m3 ) 2500


nominal diameter (m) 150
number mean diameter (m) 147.2
stan. dev. of diameter (m) 10.0

Table 2.2. Fluid flow parameters

h 20 mm
Ucl 10.5 m/s
Reh 13,800
U 0.47 m/s

Table 2.3. LDA components

Number Part Number Description


1 Lexel Model 95 4W Argon-Ion laser
2 Melles Griot 02MFG001 small 45o mirror
3 TSI 9178-2 Rotating mount/polarizer
4 TSI 9115-1 Beamsplitter
5-7 TSI 9189 3.75X Beam Expander
8 TSI 9169-450 450 mm f.d. lens
9 200 mm dia., 45o mirror
10 Dantec 9055X0341 80 mm focal length optical collector
11 TSI 9162 and 9165 Photomultiplier and Power supply
12 TSI IFA 655 Digital burst correlator
13 InfoGold P100 100 MHz Pentium lab computer
20

Figure 2.1. Particle-laden wind tunnel schematic


Chapter 2. Experimental Facility and Techniques 21

Figure 2.2. Schematic of flowmeter (from Fessler and Eaton, 1995)


22

from
particle
hoppers

tunnel cross-section

overflow drainage

Figure 2.3. Particle-feeder system (from Paris and Eaton, 2001)

x3

x2
31.1
x1

63.5

glass
windows

45.7

Figure 2.4. Test section schematic, in cm (from Paris and Eaton, 2001)
Chapter 2. Experimental Facility and Techniques 23

Figure 2.5. Magnified Image of 150 m glass particles

Figure 2.6. LDA apparatus


24

1.0

0.8

0.6
U1 / U1 CL

0.4
Paris
Kulick
0.2 Smooth Wall
Rough Wall

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

x2/h

Figure 2.7. Single phase velocity measurements


0.160
Paris
Kulick
0.120 Smooth Wall
Rough Wall
U1 rms/ U1 CL

0.080

0.040

0.000
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

x2/h

Figure 2.8. RMS velocity measurements


Chapter 2. Experimental Facility and Techniques 25

0.6

B = 0.1 (Paris)
rough wall
0.4
PDF

0.2

0.0
5 6 7 8 9 10 11 12

streamwise velocity (m/s)

Figure 2.9. Particle velocity PDF

0.8 Paris
Smooth Wall
0.6
DP/(0.5rfU1 2)

0.4

0.2

0
-250 -200 -150 -100 -50 0

x1/h

Figure 2.10. Streamwise mean pressure gradient


26
Chapter 3

Velocity Measurements in the Wall Normal Plane

This chapter presents the experimental results from this research. It is divided into two
components, gas phase statistics and particle velocity measurements for the diering wall
roughness regimes. Profile data are normalized using the mean streamwise centerline gas
phase velocity of 10.5 m/s. Where appropriate, selected data from Paris and Eaton (2001)
are presented along with the present data for comparison. The particle board apparatus
used by Kulick, Fessler, and Paris can be expected to have a roughness level roughly cor-
responding to the 150 m glass particles. This is attributed to the many years of research
use in which glass was the prevalent particulate phase utilized. The interior of the duct was
coated with glass particles. As such, that earlier case should represent a condition that is
rougher than the present smooth wall, but not as rough as the mesh screen rough wall.

3.1 Single Phase Measurement Results


Wall-normal profiles of single-phase velocities were obtained for both the rough wall and
the smooth wall cases. In each case, as with all profiles contained herein, data were taken
at eight locations in this wall normal direction. The spacing between points is reduced
as the wall is approached to more clearly show the gradients found in this region of the
flow. Only the streamwise velocity component was measured. Each datum represents an
ensemble average of four experiments, each with 20,000 individual velocity measurements.
The particles used as flow tracers were polydisperse fog oil particles placed in the plenum
chamber at the beginning of the tunnel system. As such, they pass through the blower
apparatus, which may serve to reduce their size from the manufacturer asserted nominal
0.25 m diameter.

3.1.1 Velocity PDFs


Figure 3.1 shows a typical centerline gas phase velocity PDF for both the rough and
smooth walls. The mean velocities in each of the three cases are 10.5 m/s at this location.
The present PDFs only match the most general characteristics of the PIV data obtained by
Paris (2001). As this represents the ideal case with no particle loading, it is quite interesting
to note the variations within the PDFs between the particle board case used by Paris and
previous researchers, and the smooth and rough wall cases presented herein. The deviation
between data sets, specifically the width of the PDFs and slightly higher median values in
the current cases, manifests itself in terms of higher RMS velocity values for the smooth
and rough wall cases. One explanation for this apparent dierence may very well be the
function of the dierent sizes of measurement volumes used in the experiments. In the LDA
setup used, the elliptical cross-sectional area was 75x500 m2 . In the PIV data taken by

27
28

Paris, a larger 330x330 m2 area of the channel centerline was interrogated. Because of
the larger area in the Paris measurements, all turbulence fluctuations smaller than this
length scale cannot be captured by this technique Paris (2001). While this dierence is
likely to have only minimal eects, it does highlight one potential result of the dierent
measurement techniques.

3.1.2 Mean Velocity Profiles


The mean velocity profiles for the smooth and rough wall cases are presented in figure
3.2. Excellent agreement between the data can be seen throughout the profile, indicating
that the wall boundary condition has limited eects on the gas phase mean velocities in
single phase channel flow. This result is well explained in the literature, as mentioned
in greater detail in Chapter 1, where rough wall eects on mean velocity measurements
extended only about 5 roughness element heights away from the surface (Raupach et al.
1991). With roughness elements of heights of 0.25 mm, the five-diameter rule of thumb just
reaches into the 1 mm measurement location on the measured profile, where the two data
points are nearly identical.

3.1.3 RMS Velocity Profiles


By contrast to the mean velocity profile, the wall roughness does have a significant eect
on the RMS velocity profiles. In figure 3.3 the highest turbulence levels are seen nearest the
channel wall for each of the three cases, as might be expected. It is also clear that the profiles
are the most dissimilar in the region corresponding to the channel centerline. Using only
the first three measurement locations corresponding to 0.4 < x2 /h > 1.0, this dierence
between the smooth wall and the rough wall case averages 12%. With an experimental
statistical uncertainty of 2.5% for profile data, there is clearly a marked impact of the wall
roughness elements on the turbulence levels in the center of the channel.
One possible explanation for this marked dierence at the channel centerline stems from
the work of Sommerfeld and Huber, who found a stronger correlation for the eects of wall
roughness with smaller particles than larger ones. While the flow tracers in this experiment
should be far too small to be regarded as particles, the possibility of a like eect clearly
exists. The smaller the ratio of the element size to the characteristic size of the roughness,
the larger the eects of the roughness Sommerfeld and Huber (1999). This characteristic
might manifest itself closer to the channel centerline. Regardless, it is unlikely that this
process has significant eect on the flow tracer elements.
The gas-phase fluctuating velocities in the rough wall case match or exceed those of
the smooth wall case in all regions of the flow. This indicates the possibility of turbulence
augmentation of the gas-phase induced from the flow interaction with the wall conditions.
In a similar discussion where the particle mass loading was increased, rather than the wall
roughness, a marked contrast can be seen in the work of Paris (2001) who noted that an
increase in mass loading encouraged a decrease in the gas-phase fluctuating velocities. Wall
roughness eects cannot be adequately explained with a direct comparison to increasing
the particle mass loading in the flow as cautiously suggested in Kaftori et al. (1998).
Chapter 3. Velocity Measurements in the Wall Normal Plane 29

It is curious to note, however, that despite the extreme dierences in the wall boundary
conditions in the three cases, the data for each of the four experiments depicted has its
best agreement nearest the wall. This similarity between data in the same region where
the greatest dierences in experimental conditions exist seems to imply that the physical
processes aected by the wall condition must exhibit their impact well away from their
source. A reasonable approach might consider the development of the change through
the final development section where the dierent boundary conditions exist to see if the
turbulent fluctuations of the gas phase diuse away from the wall. Likely in the region
where the roughness begins, significant dierences at the wall are found, which then tend
towards the center of the channel as the flow develops downstream.
For convenience, a table of the data used in this section is provided in table 3.1.

3.2 Two-Phase Measurement Results


In the realm of particle-phase velocity measurements, a more detailed analysis is pre-
sented for the case of the 150 m glass particles with an average mass loading of 15%.

3.2.1 Velocity PDFs


The first evidence that wall roughness has a significant impact on the particle velocity
statistics can be seen in the centerline streamwise velocity PDF depicted in figure 3.4. In
this plot, three experiments by Paris (2001) with dierent particle mass loading are shown
along with the present smooth and rough wall cases. It is clear from this plot that roughness
has a very significant eect on the particle velocity. The Paris data show a minor eect of
mass loading. At the lowest mass loading, there is a slight bi-modal character to the PDF
- this behavior is not apparent at the highest mass loading. The present rough wall case
mimics this general bi-modal PDF shape, showing a general tendency for a few particles to
travel near the 10.5 m/s velocity of the gas phase, while most particles travel at a slower
velocity, in this case approximately 9.5 m/s. In similar fashion, the Paris data has a large
tendency toward particle velocities around 8.5 m/s, with a slight increase near the 10.5 m/s
gas velocity. The smooth wall case presents a marked departure from this behavior. First,
it shows no bi-modal behavior. Further, not only do the velocities of the smooth wall PDF
exceed those of the rough wall and Paris, the smooth wall PDF actually leads the gas phase
velocity by a reasonable fraction. This characteristic, which follows intuitive expectations,
is more clearly seen in figure 3.5, which compares the smooth wall gas phase and particle
velocity PDFs.
When investigating the nature of the PDFs closer to the wall, the dierences diminish.
In figure 3.6 the smooth wall PDF still indicates average velocities in excess of those found
in the rough wall and Paris experiments, and the PDF has a much wider distribution.
This wider distribution manifests itself in increased values of the fluctuating RMS velocity,
described in more detail later. Further, in the rough wall and Paris cases, as the dierence
between the gas phase velocity and the particle velocity decreases, the distinct dierence
between a tendency for particles to travel at one of two distinct velocities is no longer clearly
30

observable. Only one velocity is evident and the apparent bi-modal characteristic of the
centerline PDFs is no longer present.

3.2.2 Mean Velocity Profiles


Figure 3.7 shows the particle mean velocity profile for the present cases plus the Paris
data at a mass loading of 0.2. Two characteristics are readily apparent from this plot.
The first shows a clear dierence between particle velocities in the smooth and rough wall
cases throughout the profile. From this observation it is obvious that the particles in the
rough wall case move slower than those in the smooth wall case at all locations. Second,
whereas Paris (2001) observed a very flat particle velocity profile, in each of the current
cases, particle velocities decrease significantly as the wall is approached. While there are
many ways to attempt to explain this slightly dierent profile shape, the most probable
cause is to infer a combination of experimental uncertainty as well as dierences in the wall
boundary condition.
Another comparison of this dierence between the smooth and rough wall velocity pro-
files can be seen in figure 3.8. This presentation of the data presents the particle slip
velocities, defined here as the dierence between the particle velocity in two-phase flow and
the single-phase gas velocity, normalized by the centerline gas velocity. The plot compares
this slip velocity with the still-air particle terminal velocity, also normalized by the center-
line gas-phase velocity. Recall from Chapter 1 that the particle terminal velocity can be
calculated from integrating the particle equation of motion, equation 1.4. From this plot,
it is apparent that for the smooth wall case, particles move faster than the gas throughout
the region of the profile examined. In Paris (2001), particles with their very flat profile
lead the flow in the region closest to the wall, and lag the two-phase flow closer to the
centerline. Conversely, rough wall particles lead the single-phase flow only at the location
nearest the wall, and lag the flow elsewhere. Finally, note that the particle slip velocity,
with particle time constant of 0.1 seconds, matches well in the smooth wall case with the
predicted terminal velocity for non-Stokesian particles only very near the wall.
In the work of Kussin and Sommerfeld (2002), it was observed that increasing the wall
roughness in a horizontal channel flow similarly served to decrease the particle velocity
for both 100 and 190 m particles. This comparison, done for a range of mass loading
conditions, indicated that the decrease in particle velocities was fairly consistent across
the profile, in much the same way as the profile in this report showed a fairly consistent
dierence between smooth and rough cases. Furthermore, particles near the channel walls
led the gas-phase velocities, again similar to the eects seen in the slip-velocity plot of figure
3.8.

3.2.3 RMS Velocity Profiles


While there are dierences in the magnitude of the mean velocities across the channel,
these dierences generally maintain similar shape and structure. In the case of the com-
parison of the RMS velocity profiles, as seen in figure 3.9, distinctly dierent shapes can be
Chapter 3. Velocity Measurements in the Wall Normal Plane 31

seen. As has been the case through most of this discussion, the data from Paris (2001) most
closely resembles the case of the rough wall. The shape of the smooth wall particle fluctu-
ating velocity profile, however, parallels that of the gas phase fluctuating velocity profile, a
trait quite evident in figure 3.10.
For the smooth wall case, the highest particle RMS velocity occurs at the nearest mea-
surement location to the wall. Contrarily, the Paris and rough wall cases have their highest
fluctuating velocities at the channel centerplane. Probably, the particle mean velocity gra-
dient is larger near the wall in the smooth wall case because particles can remain there for
a longer time. They are not forced away from the wall by the obstructive nature of the
roughness elements. This near-wall residence time can produce large particle mean velocity
gradients, which produce relatively larger particle fluctuating velocities. It is apparent that
dierent processes are at work near the wall for the smooth and rough wall cases. The rough
wall conditions likely eject the particles towards the center of the flow rapidly, reducing the
residence time of particles near the wall, and thereby reducing the velocity gradient near
the wall. This reduced gradient exhibits itself in lower particle fluctuating velocities for the
rough case near the wall.
Also of note, the location in the profile where the gas phase and the particle fluctuating
velocities cross for the smooth and rough walls is quite similar. Figure 3.11 shows that at a
value of x2 /h of 0.2, which corresponds to a distance of 4 mm from the channel wall, there
is a change in the nature of the particle RMS velocity. Specifically, closer than 4 mm from
the wall, the smooth wall particles have higher RMS velocities than rough wall particles.
Closer to the centerline, however, the rough wall particles have fluctuating velocities that
exceed those of the smooth wall. A similar, albeit less identifiable, eect occurs for the gas
phase profiles. The rough wall gas phase has higher RMS velocities on the centerline side
of the 4 mm location. The RMS velocities for the rough and smooth wall in the gas phase
are quite similar as they get closer to the wall, with identical values at the location 1 mm
from the wall. The importance of this distance, which corresponds to 27 particle diameters
or 16 roughness heights from the wall, is not readily evident from these results. However,
even in the gas phase statistics, the data for the RMS velocities for smooth and rough wall
cases presented here begins to diverge from that of Kulick and Paris at this same region of
the channel. As a final note on this peculiar behavior, the data of Paris (2001) is presented
in figure 3.12 to indicate that indeed at this same profile location the characteristics of gas
and particle phases become significantly dierent in the wall normal velocity component.
One eect contributing to the dierences in the velocity PDFs is the likely dierence in
the particle-wall collision process. In the region nearest the wall where collisions are a near
certainty for the case of the rough wall, and presumably in the case of Paris, the narrower
PDFs, and hence smaller particle fluctuating velocities, indicate a sort of velocity mixing
eect close to the wall. In the smooth wall case where these collisions are not as certain, no
such mixing is assured, and greater particle RMS velocities are seen. As regards particle-
particle collisions, this same logic should hold, as particles rebounding from wall collisions
transport their energy across the channel through collisions with other particles. If particle-
particle collisions are important closer to the channel centerplane region, as discussed by
several authors mentioned in Chapter 1, then their eect in this experiment must be to
32

enhance particle fluctuating velocities in the rough wall and Paris cases.
Sommerfeld and Huber (1999) suggest that the eects of roughness are strongly de-
pendent on the ratio of the particle diameter to the roughness height. Referring again to
figure 3.10, it is interesting to note that in the region closest to the wall, and presumably
representing those particles that have had the most collisions with both wall and neigh-
boring particles, the data of Paris shows the lowest particle RMS velocties. In the work of
Paris, the particle size and the characteristic wall roughness size are assumed to be equal.
In the smooth wall case, the ratio of particle size to wall roughness size is largest, and in
the rough wall, this same ratio is the smallest. If there is to be a correlation, therefore, it
appears as though a particle size to wall roughness ratio nearest unity produces the greatest
eects of dispersing the particle velocities through increased collisions both with the wall
and with other particles. This is a substantially dierent conclusion than the one suggested
by Sommerfeld and Huber, which suggests that the smaller this ratio between particle and
roughness sizes, the larger the resulting eects.
When comparing the results with those in Kussin and Sommerfeld (2002), less similarity
can be seen in the RMS characteristics than in the mean velocity characteristics. Whereas in
the present report rough wall particles have higher RMS velocities in the channel centerline,
while lagging near the wall, the report by Kussin and Sommerfeld (2002) shows that the
RMS velocities are uniformly higher across the channel for the high roughness case. This
dierence may possibly be a function of either the dierent particle size or mass loading, or
even the higher tunnel velocities, generally run much faster than the 10.5 m/s used in this
report. In either case, a fundamental dierence between the two experiments remains. The
augmentation in the RMS velocity associated with the increase in wall roughness for the
case of Kussin was attributed to enhanced transverse dispersion of particles. While this is
well supported in the centerline region of the current report, the explanation near the wall
provided here, namely that increased wall collisions cause higher transverse velocities, and
particles do not remain near the wall as long, does not hold in the same fashion as before.
For convenience, a table of the data used in this section is provided in table 3.2.
Chapter 3. Velocity Measurements in the Wall Normal Plane 33

Table 3.1. Gas Phase Velocity Profile Data

x2 /h Mean Velocity m/s RMS Velocity m/s


SMOOTH WALL
1 10.514200 0.497075
0.5 9.755000 0.689250
0.4 9.462500 0.749000
0.3 9.046000 0.803500
0.2 8.565000 0.858250
0.15 8.191500 0.897750
0.1 7.810250 0.955500
0.05 6.923750 1.170000
ROUGH WALL
1 10.499500 0.546700
0.5 9.445500 0.799000
0.4 9.268500 0.820750
0.3 8.895250 0.844750
0.2 8.633250 0.867500
0.15 8.120250 0.944000
0.1 7.574000 1.004500
0.05 6.921750 1.166250
34

Table 3.2. Particle Velocity Profile Data

x2 /h Mean Velocity m/s RMS Velocity m/s


SMOOTH WALL
0.03540033 10.808343 0.7080066
0.039766545 10.201779 0.7953309
0.04230261 9.788688 0.8460522
0.045466155 9.485406 0.9093231
0.04707843 9.11589 0.9415686
0.048969585 9.04617 0.9793917
0.0538587 8.701056 1.077174
0.06944112 8.026515 1.3888224
ROUGH WALL
0.48180006 9.6360012 1.22233104
0.44002035 8.800407 1.0358649
0.42328755 8.465751 0.9919413
0.41805855 8.361171 0.9997848
0.39766545 7.953309 0.951678
0.39139065 7.827813 0.9694566
0.3942666 7.885332 1.0374336
0.3602781 7.205562 1.0468458
Chapter 3. Velocity Measurements in the Wall Normal Plane 35

1.4
1.3 Paris
1.2 Smooth Wall
1.1 Rough Wall
1.0
0.9
PDF (m-1s)

0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
8 10 12

U1 (m/s)

Figure 3.1. Comparison of gas phase streamwise velocity PDFs


1.00

0.80

0.60
U1 / U1 CL

Paris
0.40

Smooth Wall
Rough Wall
0.20

0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6

x2/h

Figure 3.2. Gas phase velocity profile comparison


36

0.16
Paris
Kulick
0.12 Smooth Wall
Rough Wall
U1 rms/ U1 CL

0.08

0.04

0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

x2/h

Figure 3.3. Gas phase rms velocity comparison


0.8
B = 0.1(Paris)
0.7
B = 0.2(Paris)
0.6 B = 0.3(Paris)
Smooth Wall
0.5 Rough Wall
PDF

0.4

0.3

0.2

0.1

0.0
6 8 10 12

V1 (m/s)

Figure 3.4. Comparison of particle phase streamwise velocity PDFs


Chapter 3. Velocity Measurements in the Wall Normal Plane 37

Smooth Wall Particle


1.0
Smooth Wall Gas

0.8

0.6
PDF

0.4

0.2

0.0
8 10 12

V1 (m/s)

Figure 3.5. PDFs of streamwise velocities


0.6
Paris 0.0 < x2/h < 0.1
0.5 Smooth Wall
Rough Wall
0.4
PDF (m-1s)

0.3

0.2

0.1

0.0
4 6 8 10 12

V1 (m/s)

Figure 3.6. PDFs of particle phase velocities near the channel wall
38

1.20

1.00

0.80
V1 / U1 CL

0.60
Paris (f = 0.2)

0.40 Smooth Wall

0.20 Rough Wall

0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

x2/h

Figure 3.7. Streamwise mean velocity profile

0.10
0.08
0.06
(V1-U1) / U1 CL

0.04

0.02
0.00 Smooth Wall

-0.02 Rough Wall


-0.04
Terminal Velocity
-0.06
-0.08
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

x2/h

Figure 3.8. Slip velocity and still-air particle terminal velocity profile
Chapter 3. Velocity Measurements in the Wall Normal Plane 39

0.20

Paris

0.15 Smooth Wall


Rough Wall
V1rms/ U1 CL

0.10

0.05

0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

x2/h

Figure 3.9. Streamwise RMS velocity profile

0.20

Smooth Wall Gas Phase


0.15 Smooth Wall Particle Phase
V1rms/ U1 CL

0.10

0.05

0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

x2/h

Figure 3.10. RMS velocity profile comparison


40

0.20
Smooth Wall Gas Phase
Rough Wall Gas Phase
0.15
Smooth Wall Particle Phase
V1rms/ U1 CL

Rough Wall Particle Phase

0.10

0.05

0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

x2/h

Figure 3.11. RMS velocity profile crossing point comparison


0.10

gas-phase
0.08
particle
V2 rms/ U1 CL

0.06

0.04

0.02

0.00
0.0 0.1 0.2 0.3 0.4 0.5 0.6

x2/h

Figure 3.12. Transverse velocity fluctuation intensity for 150 m glass particles at a mass
loading of 0.2 (From Paris)
Chapter 4

Summary and Conclusions

This chapter provides a brief summary of the results obtained, draws logical conclusions
from those results, and suggests further work that might be done to extend the base of
information presented.

4.1 Summary of Experimental Results


The primary objective of this study was to measure glass particle mean and RMS ve-
locities in a fully developed vertical channel flow. This was achieved in the presence of two
significantly dierent wall roughness regimes - a smooth acrylic wall, and a similar acrylic
wall with an overlay of stainless steel mesh screen. The mesh screen was comprised of
stainless steel wire grid with a diameter of 250 m and square spacing of 1 mm between
elements. The results from both of these cases were compared with previous researchers
who conducted tests using the same apparatus with particle board development sections.
A large portion of the original apparatus was rebuilt in the process of this work. The tall
development sections, over 5 m in height, were completely rebuilt from smooth acrylic.
Only the final of these three 1.7 m sections was modified for the imposed wall roughness
condition. The design of the ductwork from the exit of the blower to the entrance of the
downward flow section was also modified from previous work.
Measurements of the gas phase velocity were taken across the profile of the fully devel-
oped channel, and compared primarily with results presented by Paris and Eaton (2001),
as well as Kulick et al (1993). In addition, 150 m glass particles of the same manufacture
as used in the Paris study were used in the two-phase experiments with mass loadings of
15%. The gas phase mean velocity profiles obtained showed that the profiles were generally
unchanged by the nature of the wall boundary condition, with excellent agreement between
the three cases. Gas phase turbulence was higher for the present cases than for the previous
experimenters at the channel centerplane region, with a slight augmentation of turbulence
in the channel centerplane for the rough wall case, as compared to the smooth wall condi-
tion. However, for all four cases near the wall, excellent agreement between the gas phase
turbulence values can be seen.
Particle velocity PDFs were significantly dierent as a result of the wall roughness
condition. At the channel centerline, the smooth wall PDF resembles a normal distribution
with mean velocity in excess of the gas phase. The rough wall and particle board cases show
the separate influence of the gas phase as well as the particle phase, resulting in a bi-modal
PDF shape. Near wall PDFs show greater variation in the smooth wall PDF, while the
rough and particle board wall PDFs indicate that a more Gaussian distribution exists in
that region.
Particle mean velocity profiles show significant dierences at all locations across the

41
42

channel profile due to the wall roughness condition. When comparing the magnitude of the
associated dierences with the work compiled by the previous experimenters, wall roughness
clearly plays a larger role in modifying particle velocities than mass loading for the regimes
studied. Smooth wall conditions show particle velocities that lead the flow as intuition
suggests. Rough wall particles lag the flow in all regions except very near the wall. Finally,
particle slip velocities increase as the wall is approached.
Particle RMS velocities dier substantially in two ways. Nearest the wall, the smooth
case suggests there are fewer particle-mixing eects, resulting in increased velocity gradients
between particles and a higher RMS velocity. After a distance from the wall of roughly 26
particle diameters, this eect is no longer seen. The rough wall case then shows steadily
increasing RMS velocities to the channel centerplane, while the smooth wall shows steadily
decreasing RMS velocities. This crossing location has additional importance in other data
sets. It represents the location in previous work where particle transverse RMS velocities
exceed their gas phase counterpart. In addition, gas phase characteristics in the streamwise
direction also show a divergence which begins at this same location.
The very flat nature of particle velocity profiles measured in previous work was only gen-
erally seen here in the rough wall case, indicating that rough wall eects likely contributed
to certain characteristics of the data presented in the earlier report. The current smooth
wall case shows behavior that more closely resembles the gas phase characteristics.

4.2 Conclusions
The results presented in this study are straightforward and clear. The consideration
of wall roughness is an important parameter in particle-laden channel flows. It has large
eects on the particle mean and RMS velocities in fully developed channel flow at all points
within a profile of the wall-normal plane. On the other hand, wall roughness seems to have
little impact on the mean velocity of the gas phase, a characteristic most clear in the region
nearest the wall. Wall roughness also slightly augments the gas phase turbulence in the
region nearest the centerplane of wall-bounded flows, while near the wall no dierences in
RMS velocities can be identified.
As regards the nature of the mechanisms behind the wall roughness eect, the particle
slip velocity may yield important information. In the present report, it was shown that
the particle slip velocity for the smooth wall case exceeded that of the rough wall case
throughout the flow, with highest values for each case in the near wall region. The influence
of the particle Reynolds number has already been established as an important parameter
in fully-developed channel flows. The ramifications of this slip velocity behavior indicates
that particle Reynolds numbers are higher in the smooth wall case than in the rough.
However, the particle slip velocities used in this report relied on average velocities, rather
than fluctuating velocities. Regardless, the implication remains that these smooth wall
particles are further from the Stokes regime than those with the rough wall boundary
condition. Particle wake eects will increase as the particle Reynolds number leaves the
very low Reynolds numbers associated with Stokes flow.
Chapter 4. Summary and Conclusions 43

Previous studies have clearly shown that particles attenuate gas phase turbulence levels
in fully developed channel flows. Increased particle wake eects were one characteristic
shown in the study by Paris (2001) to have increased the levels of gas-phase turbulence
attenuation. The mass loading was also seen to have strong eects regarding the magnitude
of that attenuation. In this report, the wall roughness condition impacts particle velocities in
a substantially greater way than the mass loading did in the study by Paris. The likelihood
that particle-particle and particle-wall collisions are increased in the rough wall regime is
quite high. As this eect is a likely contributor to gas phase attenuation, it seems reasonable
that wall roughness eects will tend to increase the attenuation of gas-phase turbulence for
the loading regime studied herein.
This project has built on the existing foundation of fully developed channel flow knowl-
edge established by a number of previous researchers. In each of those cases, the apparatus
consisted of development sections constructed of particle board wall material coated with
an accumulation of particles from years of experimentation. The recent study of Paris
and Eaton (2001) resembles the data associated with the rough wall case in the present
study. With the exception of the particle RMS velocity very near the wall, the rough wall
case matches the particle board study results in terms of particle velocity PDF shapes and
magnitudes, and general velocity profile trends.

4.3 Recommendations for Future Research


While this report has succeeded in addressing the issue regarding the importance of
the wall roughness in terms of the particle velocities, it has also created new areas which
should be explored to further the ultimate goal of establishing parameters that allow for
the predictable modeling of fully-developed channel flows. As such, there are a number of
additional areas that can be considered for additional research.
In this report, data were collected for only one phase at any given time. At no time
were both flow tracers and particles used simultaneously. A logical next step should be to
conduct similar experiments with simultaneous measurements of both gas and particle phase
velocities. In addition, only the streamwise component was measured for the experiments
here. Thus, experiments providing two-velocity component measurements using PIV or
LDA techniques that can discriminate between flow tracer and glass particles are essential
to successfully providing additional information in this regard.
Experiments combining the eects of variations in mass loading with wall roughness
eects to more fully explore the interaction of the two parameters should be conducted.
These experiments should be completed in a manner similar to that explored by Paris (2001),
where mass loadings in the range of 10-40% were utilized. In this way, the true nature
of the importance of mass loading and wall roughness conditions can best be compared.
In addition, exploring the nature of the eects studied here using a variety of particles
with dierent Stokes numbers should also be incorporated. A continued exploration of
any correlation between the ratio of particle diameter and roughness height could then be
considered.
44

Finally, taking measurements through the wall roughness development region to see the
growth of the eects away from the wall might provide substantial information regarding the
mechanisms by which particle velocities are aected in the centerplane region. While this
will provide some optical challenges for PIV or LDA techniques, continued improvement of
existing measurement techniques may provide an avenue by which this information might
be successfully obtained.
References

Bendat, J. and A. Piersol (1971). Random Data: Analysis and Measurement Procedures.
Wiley-Interscience, New York.
Fessler, J. R. (1995). Particle-Turbulence Interaction in a Backward-Facing Step Flow.
Ph. D. thesis, Stanford University, Stanford, CA 94305. Also available as Fessler, J.R.,
Kulick, J.K., & Eaton, J.K. Report No. MD-70, Mechanical Engineering Department,
Stanford University.
Fohanno, S. and B. Oesterle (2000). Analysis of the eect of collisions on the gravita-
tional motion of large particles in a vertical duct. International Journal of Multiphase
Flow 26, 267292.
Frank, T., K. Schade, and D. Petrak (1993). Numerical simulation and experimental in-
vestigation of a gas-solid two-phase flow in a horizontal channel. International Journal
of Multiphase Flow. 19, 187198.
Kaftori, D., G. Hetsroni, and S. Banerjee (1998). The eects of particles on wall turbu-
lence. International Journal of Multiphase Flow. 24, 359386.
Krogstad, P. A., R. A. Antonia, and L. W. B. Browne (1992). Comparisons between
rough- and smooth-wall turbulent boundary layers. J. Fluid Mech. 245, 599617.
Kulick, J. D. (1993). On the Interactions Between Particles and Turbulence in a Fully-
developed Channel Flow in Air. Ph. D. thesis, Stanford University, Stanford, CA
94305. Also available as Kulick, J.K., Fessler, J.R., & Eaton, J.K. Report No. MD-66,
Mechanical Engineering Department, Stanford University.
Kussin, J. and M. Sommerfeld (2002). Experimental studies on particle behaviour and
turbulence modification in horizontal channel with dierent wall roughness. Exp. Flu-
ids 33, 143159.
Paris, A. D. (2001). Turbulence Attenuation in a Particle-laden Channel Flow. Ph. D. the-
sis, Stanford University, Stanford, CA 94305. Also available as Paris, A.D., & Eaton,
J.K. Report No. TSD-137, Mechanical Engineering Department, Stanford University.
Perry, A. E., K. L. Lim, and S. M. Henbest (1987). An experimental study of the tur-
bulence structure in smooth- and rough-wall boundary layers. J. Fluid Mech. 177,
437466.
Perry, A. E., W. H. Schofield, and P. N. Joubert (1969). Rough wall turbulent boundary
layers. J. Fluid Mech. 37, 383413.
Raupach, M. R., R. A. Antonia, and S. Rajagopalan (1991). Rough-wall turbulent bound-
ary layers. Applied Mechanics Review 44 (1), 125.
Rogers, C. B. (1989). The Interaction Between Dispersed Particles and Fluid Turbulence
in a Flat-plate Turbulent Boundary Layer in Air. Ph. D. thesis, Stanford University,
Stanford, CA 94305. Also available as Rogers, Chris B. & Eaton, J.K. Report No.
MD-52, Mechanical Engineering Department, Stanford University.

45
46

Rouson, D. (1997). A Direct Numerical Simulation of a Particle-laden Turbulent Chan-


nel Flow. Ph. D. thesis, Stanford University, Stanford, CA 94305. Also available as
Rouson, D.W.I., Abrahamson, S.D., & Eaton, J.K. Report No. TSD-101, Mechanical
Engineering Department, Stanford University.
Shiller, L. and A. Naumann (1933). Uber die grundlegenden berechungen bei der schw-
erkraftaufbereitung. Ver. Deut. Ing. 77, 318.
Smalley, R. J., S. Leonardi, R. Antonia, and P. Orlandi (2002). Reynolds stress anisotropy
of turbulent rough wall layers. Exp. Fluids 33, 3137.
Sommerfeld, M. and N. Huber (1999). Experimental analysis and modelling of particle-
wall collisions. International Journal of Multiphase Flow 25, 14571489.
Tanaka, M., Y. Maeda, and Y. Hagiwara (2002). Turbulence modification in a homoge-
neous turbulent shear flow laden with small heavy particles. International Journal of
Heat and Fluid Flow 23, 615626.
Varaskin, A., Y. Polezhaev, and A. Polyakov (2000). Eect of particle concentration on
fluctuating velocity of the disperse phase for turbulent pipe flow. International Journal
of Heat and Fluid Flow 21, 562567.
Yamamoto, Y., M. Pottho, T. Tanaka, T. Kajishima, and Y. Tsuji (2001). Large eddy
simulation of turbulent gas-particle flow in a vertical channel: eect of considering
inter-particle collisions. J. Fluid Mech. 442, 303334.

You might also like