You are on page 1of 15

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

net/publication/308995183

A brief overview of the drag laws used in the Lagrangian tracking of ash
trajectories for boiler fouling CFD models

Conference Paper · September 2016

CITATIONS READS

4 3,133

3 authors, including:

Manuel García Pérez Esa K. Vakkilainen


Lappeenranta – Lahti University of Technology LUT Lappeenranta – Lahti University of Technology LUT
13 PUBLICATIONS   163 CITATIONS    206 PUBLICATIONS   2,122 CITATIONS   

SEE PROFILE SEE PROFILE

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

Steam generation from biomass View project

Pyrolysis of some biomasses widely available in Brazil View project

All content following this page was uploaded by Manuel García Pérez on 11 October 2016.

The user has requested enhancement of the downloaded file.


IMPACTS OF FUEL QUALITY ON POWER PRODUCTION, the 26th international conference
September 19—23, 2016, Prague, Czech Republic

A brief overview of the drag laws used in the Lagrangian


tracking of ash trajectories for boiler fouling CFD models
Garcı́a Pérez, Manuela; Vakkilainen, Esaa; Hyppänen, Timoa
a Lappeenranta University of Technology, P.O. Box 20, Lappeenranta 53851, Finland
The present work is concerned with the Lagrangian discrete particle tracking often
used in CFD models for ash deposition. A brief historical review of the available
drag laws and the corrections needed to account for the Cunningham effect (i.e., the
flow slip at high Knudsen numbers) is provided. Also, a newer form of a drag law
which combines previous empirical correlations in a somewhat more consistent and
reasonable way than it is suggested in the standard user’s manuals is tested with a
simple CFD ash deposition model.
It is shown that within a boiler the mean free path of the flue gas is of the order of
0.1 µm. This value may result in corrections to the drag law (usually, the Stoke’s
particle flow regime) that could vary spatially from a 39 to a 68 % for particles of half
a micron diameter. This correction increases even more at higher Knudsen numbers
(finer particles). The validity of these high corrections is questioned.

Keywords: Fouling; Drag law; Boilers; Ash deposition; CFD modeling.

1. Introduction
The CFD modeling of disperse, discrete particle-laden fluid flows is of high interest in a
wide variety of engineering fields [1]. This approach consists of the injection and trajectory
computation of discrete particles within the CFD domain. Hence, it is often referred to as the
‘discrete phase model’, or as the ‘Lagrangian-Eulerian’ model (a Lagrangian tracking scheme of
particles within an Eulerian solving of the continuous carrier phase by means of the Navier-Stokes
equations).
The ash deposition and accumulation in cold surfaces of industrial furnaces is a particularly
active field which benefits from this tool. Numerous researchers have used this Lagrangian par-
ticle frame to model the fly ash trajectories accounting for fouling and slagging [2–9]. Typically,
little attention has been put in the accuracy of the discrete phase trajectory solving in compari-
son to the accuracy of the dense phase flow solving [5,10]. When the particles have a moderate to
high Knudsen number (Kn > 0.1), the carrier phase (the flue gas) starts to not behave perfectly
as a continuous medium, entailing a flow slip in the boundaries with solid walls and particles.
These conditions, usually referred to as ‘rarefied’ flow or ‘Knudsen’ flow, have been of interest
for high-altitude aircraft, high-vacuum and microchannel applications [11]; but unfortunately
little work has been done recently regarding the flow slip effects in the particles’ motion which
is of considerable interest for fine ash deposition in combustion environments.
A direct consequence of this rarefied flow phenomenon is that the often used finite-volume
CFD models tend to overestimate the drag that the discrete fly ash particles are exerted within
the flow. It will be highlighted how, unfortunately, the user’s guides and manuals of some CFD
packages might be somewhat incomplete and misleading regarding this issue, as inconsistent
formulations and models are recommended. In this work, a modified form of these drag laws
for particle tracking is proposed for boiler ash particulate. In addition, a simple ash deposition
CFD model inspired in previous work [8] has been implemented to examine this new approach.

1
IMPACTS OF FUEL QUALITY ON POWER PRODUCTION, the 26th international conference
September 19—23, 2016, Prague, Czech Republic

2. Background

2.1. Laws for flow drag on spherical particles


The trajectory of a particle is calculated by integrating the following force balance [12]:
d~v
mp = mp FD (~u − ~v ) + F~other , (1)
dt
where ~v is the particle velocity, ~u is the fluid velocity in the vicinity of the particle, and F~other
stands for forces of any other nature (e.g., Brownian, pseudo-random turbulent eddies, ther-
mophoresis, or gravity if required) acting on the particle. The term FD (~u −~v ) is the acceleration
that the fluid exerts to the particle. FD is typically addressed as a specific function of the particle
Reynolds number:
18µ CD Rep
FD = , (2)
ρp d2p 24
where the drag coefficient CD is itself a function of Rep as well. ρ and µ stand for the fluid
density and molecular viscosity, and dp is the particle diameter. The particle Reynolds number
is defined here as Rep = ρ|~u − ~v |dp /µ. The previous equation, including the selected formulation
for of CD , is usually referred to as ‘the drag law’.
At sufficiently low particle Reynolds number, the viscosity dominates the particles’ motion.
It is then possible under this assumption to simplify, linearize, and integrate analytically the
Navier-Stokes equations for the particle drag. The solution for a spherical particle yields an
expression for the drag coefficient as CD = 24/Rep , justifying the particular arrangement of the
previous equation. This motion regime is often referred to as Stokes’ flow, or creeping flow.
For higher values of Rep , different approaches have been reported. For instance, Schiller and
Naumann [13] suggested the following formulation for CD :

 24/Rep Rep < 0.5
0.687
CD = 24(1 + 0.15Rep )/Rep 0.5 < Rep < 1000 (3)
0.44 Rep > 1000

Similarly, Morsi and Alexander [14] proposed a more complete approach in the form of:
a2 a3
C D = a1 + + 2 , (4)
Rep Rep

with the coefficients a1 , a2 , a3 shown in table 1 for the ranges of Rep which could be of possible
interest for ash fouling and slagging in boilers.

Table 1: Coefficients a1 , a2 , a3 of the model of Morsi and Alexander for Rep ≤ 1000 [14]. Note how
the first row corresponds to Stokes’s drag law (i.e., CD = 24/Rep ).

range a1 a2 a3
Rep < 0.1 0 24 0
0.1 < Rep < 1 3.690 22.73 0.0903
1 < Rep < 10 1.222 29.1667 -3.8889
10 < Rep < 100 0.6167 46.50 -116.67
100 < Rep < 1000 0.3644 98.33 -2778

2
IMPACTS OF FUEL QUALITY ON POWER PRODUCTION, the 26th international conference
September 19—23, 2016, Prague, Czech Republic

2.2. The Cunningham correction


A particular phenomenon occurs with very small particles. The flue gas within a boiler may
not behave perfectly as a continuous medium at that reduced scale since very small particles
might need to travel a certain distance before being impacted by enough gas molecules to expe-
rience the expected drag predicted by the previous drag laws. In other words, particles that are
small enough may experience some slip within the gas molecules. The importance or strength
of this effect is represented by the particle Knudsen number Kn = λ/(dp /2), where λ stands
for the mean free path which is the average distance travelled by a moving particle between
successive impacts with gas molecules. For sake of consistency with the notation used in the
studies mentioned and referenced below, this Kn number is based on the particle radius instead
of being based on its diameter. Different scientists have reported slightly different expressions
for the determination of λ [15]. Usually, these expressions can be recast as a function of the flow
properties:
r
µ πMw
λ= , (5)
φρ 8RT
where R is the ideal gas constant, MW is the gas molecular weight, T is the fluid temperature
and the value of φ is 0.3502 according to Boltzmann’s kinetic-molecular gas theory [16] or 0.491
according to Chapman-Enskog’s theory [17] (somewhat more elaborated than Boltzmann’s).
Hence, the Knudsen number compares the particle size to the distance that it travels before
being affected by gas molecules. By using the expression of Chapman-Enskog, the mean free
path of the ambient air is about 0.07 µm. The mean free path of the flue gases of boilers could
reach values of about 0.14 µm (at 300 o C) and 0.34 µm (at 900 o C). These lengths are definitely
comparable to the sizes of the fine fly ashes (Kn ∼ 1). Logically, the larger the Kn, the more
relevant the particle slip within the gas becomes. As it will be seen with the following equations,
the slip effect may decrease the drag by a 11.5% already at about Kn ≈ 0.1.
Cunningham [18] addressed this deviation by introducing a correction factor Cc > 1 which
would be dividing the drag, in order to account for the slip. From kinetic considerations on
the statistical amount and strength of the impacts that a particle receives by the surrounding
gas molecules, he concluded that this correction factor would depend on the on the Knudsen
number as Cc = 1 + Kn · A, with A a parameter which depends on the nature of the gas.
Later, Millikan [19] confirmed this phenomenon empirically and implemented this correction
to calculate accurately the gas drag on the small oil drops of his noted experiment where the
charge of the electron was measured. Cunningham and Millikan proposed respective values for
A (adapted from the Boltzmann’s definition of λ used in their work to the definition found in
the Chapman-Enskog theory) of 1.105 and 1.146.
Davies [20] reviewed other correlations made by other authors for Cc factor which treated
A as an exponential function of the inverse of Kn. Based on the parameters obtained by other
researches, he proposed the following expression for air:

Cc = 1 + Kn [1.257 + 0.400exp(−1.1/Kn)] . (6)

Alternatively, Allen and Raabe [15] proposed a slightly more complicated correction factor found
by the best fit to the data reported and tabulated by Millikan, with the aim to find a gppd
approximation to the experimental measurements especially at moderately low Kn numbers:
    
−0.0026 −0.74
Cc = 1 + Kn 1.205exp + 0.425exp . (7)
Kn Kn

3
IMPACTS OF FUEL QUALITY ON POWER PRODUCTION, the 26th international conference
September 19—23, 2016, Prague, Czech Republic

All these coefficients in these correlations given here correspond to the Chapman-Enskog
formulation of the mean free path (and consequently, the Knudsen number). These correlations
are plotted in Fig. 1, where it can be seen that they yield approximately similar values at the
Kn numbers of interest for boiler applications. Their behavior is linear or almost linear.

Cc
3

1
0.01 0.1 1
Kn
Cunningham Millikan
Davies Allen and Raabe

Figure 1: Cunningham correction factor proposed by Cunningham [18], Millikan [19], Davies [20] and
Allen and Raabe [15]. These charts correspond to the Chapman-Enskog definition of λ.

This Cunningham correction has been discussed and studied for small drops and particles,
usually in terminal fall within a fluid following the aforementioned Stokes’ flow regime (Rep <
0.1), at very low Reynolds numbers. Indeed, the Cunningham effect has often been mentioned
as a deviation from the Stoke’s law. Little or no work has been done on the possible effects of the
particle slip within the fluid at larger Reynolds numbers. It is possible, due to the flow features
of boilers, to find some ash particles outside of the Stoke’s regime (Rep > 1) and still showing
moderate to high Knudsen values (Kn > 0.1). It is logical to think that similar slip conditions
might take place in other Reynolds regimes. Due to the lack of a properly documented empirical
work on this issue, a reasonable approach would be to apply the Cunningham correction to the
Reynolds ranges beyond the Stoke’s regime as well. Nonetheless, the largest Reynolds numbers
in ash particles are typically (but not always) a consequence of their larger sizes, which would
ultimately have a low Knudsen number and consequently require low Cunningham corrections.

2.3. Drag laws in Fluent and OpenFOAM


Different CFD packages offer and/or suggest different implementations by default for the
drag law. OpenFOAM documentation suggests a slightly modified version of the law of Schiller
and Naumann [13]. Ansys Fluent [12] suggests also the Schiller and Naumann law as ‘generally
acceptable’ for solid-fluid Eulerian-Eulerian multiphase simulations, but, in the other hand, the
Morsi and Alexander [14] law is recommended and available by default for the discrete Lagrangian
particle tracking. In addition, as general practice it is suggested to use ‘for submicron particles’
the Stokes’ flow regime accounting for the Cunningham correction, suggesting the expression
provided by Davies [20] (Eq. 6) with the coefficients adapted to the Boltzmann’s definition of
the mean free path. Unfortunately, this is done without specifying at all the mean free path
calculation, and without referencing Davies’ equation properly. These instructions are misleading
and may lead to the computation of an erroneous Cc [6] if these guidelines are observed without
a deeper investigation as the one that has just been provided in section 2.2.
It will be noted in section 4 that particles greater than a micron (e.g., dp = 3 µm) might
still experience a significant slip from the Stokes’ drag regime of around a 10.9%. Indeed, there

4
IMPACTS OF FUEL QUALITY ON POWER PRODUCTION, the 26th international conference
September 19—23, 2016, Prague, Czech Republic

is no specification in Cunningham’s and Millikan’s work suggesting that this correction is to be


applied for submicron particles only and not for larger particles; yet such a guideline appears
indicated in the Fluent User’s theory guides [12]. Moreover, when the default in-built Stokes-
Cunningham drag law is selected, the user is required to input just a value for the Cunningham
correction which would be applied to particles of all diameters and regardless of their location
and of their Knudsen number, which does vary spatially.
Although it is possible to code and to implement a user-defined customized drag law to
circumvent this (with the UDF routine DEFINE DPM DRAG [21]), Ansys Fluent (up to release
16.2) disables the simulations of the Brownian motion and the turbulent eddy effects on particle
trajectories when the default (and insufficient) in-built Stokes-Cunningham drag law is not in
use. It might be possible to implement the Brownian motion separately by using a different
UDF routine (DEFINE DPM BODY FORCE), which would stand for the term F~other of Eq.
1. Unfortunately, the discrete random walk model used to simulate the turbulent dispersion of
particles could not be programmed manually with the available code and documentation, since
it requires the per-parcel storage and usage of the eddy remaining lifetime and length. These
variables are generated separately for each parcel and they are maintained until the tracked
particle leaves the eddy or the eddy fades; then a new eddy would be generated according to
the turbulence intensity of the particle vicinity [12]. The variables or references where these
eddy parameters are stored or generated are not documented, and it is not reliable to establish
separate variables for this purpose outside the particle data structure since the routine is passed
a pointer to the particle data structure itself and not any related external variables at the user’s
will. The effects of the turbulence on the particle trajectories are significantly relevant to the ash
particle diffusion and trajectories, having important consequences regarding their deposition in
surfaces when pseudo-random eddies project the particle through the tube boundary layer [22].
Fortunately, a newer Fluent release (version 17.0) allows for the usage of a customized drag
law along with the random-walk simulation of the turbulence effects. Therefore it is possible now
to account for all the necessary phenomena by only modeling the Brownian motion separately
(DEFINE DPM BODY FORCE). A component i of a Brownian force is modeled as [12]:
s
6πµkB T dp
FB,i = ζi , (8)
∆tp Cc

where ζi is a zero-mean unit-variance independent Gaussian pseudo-random number generated


at each particle time-step, kB is the Boltzmann’s constant (1.381·10−23 J/K) and ∆tp is the
time-step used for the integration of the particle trajectory (this is, typically, the flow time-step
but it could be chosen differently at will).

3. A modeling case study


In order to highlight the relevance of the selection of the drag law, and also, the possible
importance of calculating spatially the Cunningham correction each time step for each particle;
a simple two-dimensional unsteady CFD model has been developed. The purpose of this study
is to clarify the effects and consequences of specific drag laws, and not to perform a deep analysis
on fouling phenomena which could be found elsewhere in some of the references already provided
( [3,5,7,8]). Therefore the model features will be kept reasonably simple. The ash deposition on
a water-cooled single-tube probe (4 cm diameter, surface temperature 39 o C) within a PC-fired
boiler backpass (upstream temperature 600o C, upstream fluid velocity 4 m/s, pressure 92 kPa)

5
IMPACTS OF FUEL QUALITY ON POWER PRODUCTION, the 26th international conference
September 19—23, 2016, Prague, Czech Republic

is simulated. Other flue gas properties have been selected consistently with the temperature and
pressure according to [23]: specific heat 1248 J/(kgo C), thermal conductivity 0.0663 W/(m·o C),
viscosity µ = 3.82 · 10−5 kg/(m·s) and molecular weight Mw = 28.97 g/mol.

3.1. Mesh and boundary conditions


The two-dimensional domain consists of a rectangle of 20 cm height and 52 cm long. The
left side of the rectangle will be the velocity inlet of the gas at 4 m/s and the opposite side is the
pressure outlet. The remaining sides are set as periodical boundaries. A hole of 4 cm diameter
is placed at 18 cm downstream of the inlet representing the water-cooled probe.
The domain has been meshed with a pave of triangular cells. The perimeter of the probe
has been first divided in 700 elements in order to account for numeric accuracy guidelines stated
in a previous study [5]. Afterwards, a size function was set to control that the size of the cells
increase with a ratio of 1.25 up to a maximum cell area of 30 mm2 . The resulting grid contains
30110 cells and can be seen in Fig. 2.

Figure 2: Domain mesh.

3.2. Solver and procedure


The model is executed in Ansys Fluent 17. The single-precision solver with default discretiza-
tion schemes are used. The selected time-step of 5·10−5 s allows for a maximum Courant number
around 2, satisfying the time-resolution conditions of [24]. A limit of 40 iterations per time-step
allows the normalized residuals to stabilize before solving the following step. A viscous SST k-ω
turbulence model with standard input was chosen.
The flow past transversal cylinders or tubes experiences periodical unsteady oscillations and
vortex shedding [25], justifying the need for an unsteady simulation. Once reached quasi-stable
flow patterns in this model, the period of the oscillation was 41.97 ms. For each one of the
two simulations which are carried out (with different drag law schemes), the calculations will be
executed during an exact duration of 20 of these periods (in flow time) to reduce the potential
result bias which could occur with a too short simulation time. During that flow duration the
relevant parameters of each impaction are logged for posterior data postprocessing. This strategy
has been already utilized in another approach [8].

3.3. Coal ash and size distributions


Coal ash particles will be simulated. The ash properties and size distributions of a field study
on a pulverized coal combustor [26] will be used. The ash particles have a density of 3170 kg/m3 ,

6
IMPACTS OF FUEL QUALITY ON POWER PRODUCTION, the 26th international conference
September 19—23, 2016, Prague, Czech Republic

a thermal conductivity of 11.26 W/(m·o C) and a specific heat of 796 J/(kg·o C).
The measured particle size distributions are approximated here with the superposition of
three separate Rosin-Rammler distributions. The total range of particle sizes is therefore divided
in three intervals or subranges. Each one of those intervals is assigned separate and independent
Rosin-Rammler distribution parameters. A Rosin-Rammler particle distribution is modeled as
  n 
dp
Yd (dp ) = exp − , (9)
d
where Yd (dp ) is the fraction of particles greater than dp , d is the mean particle diameter and the
parameter n is called the ‘spread factor’ [24]. A distribution is thus completely determined with
d, n, and the minimum and maximum particle diameters. Table 2 states the parameters of the
three injections utilized in this study which make the best adjustments to the measurements of
the previous empirical work [26].

Table 2: Parameters of the three used Rosin-Rammler injections.

Distribution \Subrange Fine Middle Coarse


min. dp [µm] 0.0146 0.0615 1.11
d [µm] 0.0306 0.873 3.391
max dp [µm] 0.0533 0.9650 14.86
n 3.32 3.19 1.65
concentration in the gas [mg/m3 ] 1.739 29.63 3745
flow at inlet [g/s] 0.00139 0.0237 2.996

The injections are modeled by discretizing each one of them into 10 logarithmically-distributed
different diameters (between the minimum and the maximum sizes). At each time step, a parcel
of particles of each one of these sizes is released through the inlet boundary.

3.4. Drag laws


The main contribution of this work is to highlight the new drag law proposed, which is a
combination of the Morsi and Alexander drag with the Cunningham correction:

18µ CD Rep
FD = , (10)
ρp d2p 24Cc
with CD is given by Eq. 4 and Table 2 [14]; and the expression of Allen and Raabe [15] (Eq.
7) has been chosen for Cc . The mean free path for the Cunningham correction is calculated for
each particle at each time step as a function of the pressure and temperature of the gas cell
containing the particle (Eq. 5). This customized drag law must be coded by a user-defined
function since it is not available as such by default in the CFD software package. This approach
has been used in a recent ash deposition model enhanced with dynamic meshes [9].
In order to contrast these results with the default available options, a second simulation is
carried out where the only difference is the drag law. As suggested by the manuals, the default
Morsi and Alexander law will be used for the coarse injection (dp > 1 µm) while the Stokes-
Cunningham drag is used for the middle and fine injections instead. It is not possible to compute
Cc individually for each particle, and the user is prompted to introduce a constant and uniform
value for each injection. It is uncertain whether it is most reasonable to compute λ with the inlet

7
IMPACTS OF FUEL QUALITY ON POWER PRODUCTION, the 26th international conference
September 19—23, 2016, Prague, Czech Republic

flue gas conditions (since these conditions are present within most of the domain) or to take some
conditions within the tube boundary layer (since our end target is the ash deposition); remarking
more the important limitations of this default approach. For this case, we have selected the inlet
conditions, which imply that λ = 0.2726 µm. By using the mean diameters of the fine and
middle distributions, we obtain respective Kn numbers of 17.8 and 0.62; which have respective
Cunningham corrections Cc of 22.4 and 1.92.
The simulation of thermophoresis, random eddy lifetime, and Saffman lift forces were enable
for this work. The Brownian motion was also implemented separately with a customized user-
defined function according to Eq. 8.

3.5. Sticking model


A mechanistic particle sticking-rebound submodel is implemented in order to calculate the
particle adhesion rates and the rebound velocities if a particle does not stick; with the aim to
provide the end results of the drag laws tested in this study. The approach used in this work is
an adaptation of two separate sticking criteria.
A model developed by van Beek [27] considers that an impacting particle converts its incident
kinetic energy into stored mechanical energy at the moment of contact, when its velocity is zero
and the deformation is maximum. If this stored energy Estored is large enough to create new
surface (with surface energy Es ) and outgoing kinetic energy Ek , the particle detaches. Therefore
the sticking criterion of van Beek can be summed up as Estored < Es + Ek .
On the other hand, Konstandopoulos [28] stated that no particle may stick onto a powdery
layer if the incidence angle is too oblique, regardless of the energy criterion mentioned above.
There exists a certain critical incidence angle θcr , which is a function of the surface friction
coefficient and the material Poisson’s ratio, beyond which an incident particle would always
rebound. Therefore, the sticking criterion of Konstandopoulos may be summed up as θ < θcr .
The sticking/rebound submodel used here combines those two criteria, which must be satis-
fied simultaneously for a particle to stick. This approach has also been used by other researchers
in recent studies [8, 29, 30]. The implementation and calculation used here is identical to the
one performed in previous research by the same authors of this study [8], where the submodel
is explained in more depth. A complete explanation would be long, requiring several pages and
are thus omitted since they are available elsewhere [8, 27, 28] and the sticking submodel is not
the key feature of this study, but just a means to obtain the end results.
The mechanical properties of the ashes regarding stick/rebound have been taken as the ones
of SiO2 , which was the major component in the ash chemical analysis [26]. These properties
are: friction coefficient 0.7 [27], Young modulus 3 · 1010 Pa [31], Poisson’s ratio 0.3 [31], surface
energy SiO2 —gas 0.15 J/m2 [32]. The Konstandopoulos critic angle is θcr = 76.48o .

4. Results and discussion

4.1. Spatial variations of the corrections


Fig. 3 highlights the instantaneous field of the mean free path, the Knudsen number and
the Cunningham correction factor for particles of 0.5 µm. It can be noted how the correction
varies spatially from a 39% (in the areas immediately downstream of the probe) to a 68% (in the
wake and in the free-shear regions). It is of most importance to calculate the particle trajectory
reliably in all areas for ash deposition prediction, especially when modeling tube arrays [8]. Such

8
IMPACTS OF FUEL QUALITY ON POWER PRODUCTION, the 26th international conference
September 19—23, 2016, Prague, Czech Republic

a spatial consideration of the Cunningham correction is not possible with the default in-built
drag laws, and must be thus customized.

Figure 3: Field of the mean free path (upper scale), Knudsen number for particles of 0.5 µm diameter
(middle scale), and their respective Cunningham correction factor (bottom scale).

The range of Knudsen numbers of interest here is narrow enough so that the behavior of
the Allen and Raabe correlation is nearly linear (compare in Fig. 1 with the linear lines of
Cunningham and Millikan) even though the expression has exponential terms (Eq. 7). Therefore,
the same field represented in Fig. 3 may be also used to understand the flow slip that particles of
others diameters experience by just using different scales for Kn and Cc . Particles of 3 µm have
Knudsen numbers ranging from 0.0513 (in the rear vicinity of the probe) to 0.0867 (in the shear-
free flow region) with respective Cunningham corrections of 6.4 % and 10.9 %. Consequently,
particles that are greater than a micron still may be affected by the gas slip as well. This
phenomena may not be limited to submicron particles as it is pointed out in the manuals.
It should be mentioned as well what occurs to significantly smaller particles. Particles of e.g.
0.05 µm will have Knudsen numbers as high as Kn = 5.2 in the free-shear regions, which result
into Cunningham corrections of Cc = 9.2, about a 820% of deviation from the uncorrected drag
law. It is intuitive to think that an equation or law which has to be corrected by a 820% is very
likely to be working out of its range of validity, even though it was derived analytically [18].
Empirical work must be performed within this range of very high Knudsen numbers (Kn > 1)
with the aim to determine the reliability of the Cunningham-corrected creeping flow theory.
For instance, some studies start to account for corrections in the molecular viscosity and the
introduction of flow slip coefficients near the wall areas to still solve their rarefied flows by using
the standard continuous versions of the Navier-Stokes equations for fluids in microchannels [11].
Unfortunately, little work of this kind has been done for the particles’ drag at moderate to
high Knudsen numbers. Such a study, although it would be of an immense scientific interest,
falls however outside of the scope of the present work. Researchers are therefore encouraged to
interpret with great care the results obtained for particles at already Kn > 0.1.

9
IMPACTS OF FUEL QUALITY ON POWER PRODUCTION, the 26th international conference
September 19—23, 2016, Prague, Czech Republic

4.2. Diameter-wise sticking deposition magnitudes


Fig. 4 shows the relevant deposition magnitudes as a function of the particle diameter which
result from the application of both drag laws (the standard one referred to as ‘old’, and the law
proposed here referred to as ‘modified’). The arrival rate accounts for the amount of mass which
travel and impact onto the probe surface, without taking into account whether they stick or
not. Therefore, the simple relationship (deposition rates) = (arrival rates) · (sticking efficiency)
applies for each dp [5, 7].

100 1.2

80 1
0.8
60

mg/s
0.6
%

40
0.4
20 0.2
0 0
0.01 0.1 1 10 0.01 0.1 1 10
dp [mm] dp [mm]
(a) (b)
0.025 0.01

0.02 0.008

0.015 0.006

0.01 0.004

0.005 0.002

0 0
0.01 0.1 1 10 0.01 0.1 1 10
dp [mm] dp [mm]

(c) (d)

Figure 4: Relevant deposition magnitudes as a function of the particle diameter dp . (a): Sticking
efficiencies. (b): Deposition rates on the probe. (c): Normalized arrival rates. (d): Normalized
deposition rates.

In addition, this work makes use of ‘Normalized’ deposition magnitudes in a similar fashion as
it has been done in a previous approach [8]. The normalized deposition rates and the normalized
arrival rates are independent of any inlet the ash particle size distribution or concentration and
allows for a better comparison of the behaviors of the particles of different sizes, regardless
of how much share of those particle diameters is present in the flue gas. It can be noted
(Table 2) that, for instance, the finest distribution has a total mass fraction of about 3000 times
less than the coarse distribution. It is thus expected to see very small contributions to the

10
IMPACTS OF FUEL QUALITY ON POWER PRODUCTION, the 26th international conference
September 19—23, 2016, Prague, Czech Republic

total deposition rates from these fine particles compared to the largest particles. The simple
relationship: (normalized magnitude) = (absolute magnitude, in kg/s) / (mass flow input for
each diameter) is used to build the normalized rates and it allows for a better analysis of the
behavior of the particles as a function of dp . The resulting normalized rates are dimensionless
and they are shown in the bottom half of Fig. 4.
It can be noted how, although it is not appreciable in the total deposition rate -chart (b)-
the traditional drag law presents remarked discontinuities in the transitions between different
Rosin-Rammler ranges (which occur at dp =0.6 µm and at dp = 1.0 µm). Such a discontinuity
does not occur with the modified drag law proposed here, as a smooth trend for all magnitudes
is appreciated up to diameters of approximately 1—2 µm. The sticking efficiency decreases
because the bigger particles possess a greater mass and thus a greater kinetic energy, making it
more difficult to stick according to the van Beek energy criterion [27].
The normalized arrival rates were practically constant and uniform for particles up to 3 µm.
The Stokes number of these 3-micron particles, calculated as:

ρp d2p |~v∞ |
Stk = (11)
9µDtube

is 8.3 · 10−3 . For any Stk below that value, the particles showed identical normalized arrival
rates, suggesting that they are not prone to deposit by inertia anymore and thus other deposition
mechanisms are taking place. From this diameter, there is a remarked steep increase in both
the arrival and deposition rates, meaning that the inertial impaction mechanism starts to affect
suddenly beyond this threshold only.
The slightly descending trend shown in the normalized deposition rate -chart (d)- is explained
by a similar slightly descending trend observed in the sticking efficiency. The pronounced drop
of the efficiency at dp > 3 µm does not compensate for the large fraction of particles that arrives
with inertia and hence the deposition rate increases suddenly.

4.3. Ash deposition around the probe


Fig. 5 (left) shows the aggregated deposition over the probe perimeter. It can be noted
how the distributions practically coincide regardless of the drag law. This is a consequence, as
mentioned before, of the relatively large fraction of big particles, which respond similarly to both
drag law schemes. However, the deposition rates for the smallest particles (Fig. 5, right) show
a significant discrepancy between both schemes, as the default one overestimated the deposition
by a 36.6 % (in average).
As a consequence it could be concluded that when the small particles represent an important
share of the ash load (e.g., in economizers of kraft recovery boilers where the fume particles
created by alkali vapor condensation represent the majority of the incoming ash in the flue
gas [22, 33]), these improved drag laws should be considered (e.g., it was done in [9]). On the
other hand, the default and simpler drag law schemes could be reliably used instead only if the
intermediate to large particle sizes (dp > 5 µm) are relevant in the diameter distributions and
the corresponding Knudsen numbers are mainly below 0.1.

5. Conclusions
This work was concerned with the decrease of the drag that fine fly ash particles may experi-
ence within the flue gas as a consequence of the flow slip that results from the mean free path of

11
IMPACTS OF FUEL QUALITY ON POWER PRODUCTION, the 26th international conference
September 19—23, 2016, Prague, Czech Republic

0.16 0.0016

0.12 0.0012

g/m2s
g/m2s

0.08 (a) 0.0008 (b)

0.04 0.0004

0 0
-1 -0.5 0 0.5 1 -1 -0.5 0 0.5 1
a/p a/p

Figure 5: Simulated deposition rates over the tube perimeter. α is the angular coordinate of the probe,
and it is set so that α = 0 at the tube lee and α = ±π at the tube wind. Left: aggregated deposition of
all particles. Right: deposition of particles smaller than 0.8 µm.

the gas being comparable to the size of the particle. A review on how this problem was addressed
in literature has been provided. It has been made patent that a remarkably little amount of
work has been performed on this phenomenon regarding industrial boiler ash applications, which
would be of great value for deposition models.
Although there is current research going on other applications of Knudsen flows, most of
the relevant work done for particle-laden fluids date from the last century and were limited to
correlate the Cunningham correction. Additional work is encouraged here for a proper statement
of how the Knudsen flow conditions affect the particle trajectories, and for empirical confirmation
of these corrections when Kn > 0.5 (which may happen easily in industrial furnaces for the
finest ash particles). This condition entails very high values for the Cunningham correction.
Consequently, under these circumstances it is intuitive to think that its validity may certainly be
compromised, even though their usage appears recommended in manuals and literature without
specifying any limitations.
A CFD model was implemented to obtain a qualitative idea of the importance of this issue
concerning the coal ash fouling on a deposition probe. The spatial variation of the Cunningham
correction, which also depends on the particle diameter, has been highlighted. Although the
difference of the effects of the two drag law schemes were negligible for big particles (dp > 3
µm), the finer particles behaved differently. This should be taken into account when modeling
environments with important loads of fine dust like e.g. fume ash in kraft recovery boilers.

References

[1] S. I. A. for Multiphase Flows (SIAMUF) and M. Sommerfeld, Best Practice Guidelines
for Computational Fluid Dynamics of Dispersed Multi-Phase Flows. European Research
Community on Flow, Turbulence and Combustion (ERCOFTAC), 2008.

12
IMPACTS OF FUEL QUALITY ON POWER PRODUCTION, the 26th international conference
September 19—23, 2016, Prague, Czech Republic

[2] J. Tomeczek and K. Waclawiak, “Two-dimensional modelling of deposits formation on


platen superheaters in pulverized coal boilers,” Fuel, vol. 88, no. 8, pp. 1466–1471, 2009.
[3] H. B. Vuthaluru and R. Vuthaluru, “Control of ash related problems in a large scale tan-
gentially fired boiler using CFD modelling,” Applied Energy, vol. 87, no. 4, pp. 1418–1426,
2010.
[4] R. Weber, M. Mancini, N. Schaffel-Mancini, and T. Kupka, “On predicting the ash be-
haviour using computational fluid dynamics,” Fuel Processing Technology, vol. 105, pp. 113–
128, 2013.
[5] R. Weber, N. Schaffel-Mancini, M. Mancini, and T. Kupka, “Fly ash deposition modelling:
Requirements for accurate predictions of particle impaction on tubes using RANS-based
computational fluid dynamics,” Fuel, vol. 108, pp. 586–596, 2013.
[6] M. Garcı́a Pérez, E. K. Vakkilainen, and T. Hyppänen, “2D dynamic mesh model for
deposit shape prediction in boiler banks of recovery boilers with different tube spacing
arrangements,” Fuel, vol. 158, pp. 139–151, 2015.
[7] A. M. Beckmann, M. Mancini, R. Weber, S. Seebold, and M. Müller, “Measurements and
CFD modeling of a pulverized coal flame with emphasis on ash deposition,” Fuel, vol. 167,
pp. 168–179, 2016.
[8] M. Garcı́a Pérez, E. Vakkilainen, and T. Hyppänen, “Unsteady CFD analysis of kraft recov-
ery boiler fly-ash trajectories, sticking efficiencies and deposition rates with a mechanistic
particle rebound-stick model,” Fuel, vol. 181, pp. 408–420, 2016.
[9] M. Garcı́a Pérez, E. Vakkilainen, and T. Hyppänen, “Fouling growth modeling of kraft re-
covery boiler fume ash deposits with dynamic meshes and a mechanistic sticking approach,”
Fuel, vol. 185, pp. 872–885, 2016.
[10] F. Greifzu, C. Kratzsch, T. Forgber, F. Lindner, and R. Schwarze, “Assessment of particle-
tracking models for dispersed particle-laden flows implemented in OpenFOAM and ANSYS
FLUENT,” Engineering Applications of Computational Fluid Mechanics, vol. 10, no. 1,
pp. 30–43, 2016.
[11] N. Dongari and A. Agrawal, “Modeling of navier–stokes equations for high knudsen number
gas flows,” International Journal of Heat and Mass Transfer, vol. 55, no. 15, pp. 4352–4358,
2012.
[12] ANSYS Inc. (US), “ANSYS Fluent 17.0 Theory Guide,” 2016.
[13] L. Schiller and Z. Naumann, “A drag coefficient correlation,” Vdi Zeitung, vol. 77, no. 318,
p. 51, 1935.
[14] S. Morsi and A. Alexander, “An investigation of particle trajectories in two-phase flow
systems,” Journal of Fluid Mechanics, vol. 55, no. 2, pp. 193–208, 1972.
[15] M. D. Allen and O. G. Raabe, “Re-evaluation of millikan’s oil drop data for the motion of
small particles in air,” Journal of Aerosol Science, vol. 13, no. 6, pp. 537–547, 1982.
[16] L. Boltzmann and S. G. Brush, Vorlesungen über Gastheorie, vol. 1. Akademische Druck-u.
Verlagsanstalt, 1896.

13
IMPACTS OF FUEL QUALITY ON POWER PRODUCTION, the 26th international conference
September 19—23, 2016, Prague, Czech Republic

[17] S. Chapman and T. G. Cowling, The mathematical theory of non-uniform gases: an account
of the kinetic theory of viscosity, thermal conduction and diffusion in gases. Cambridge
university press, 1939.
[18] E. Cunningham, “On the Velocity of Steady Fall of Spherical Particles through Fluid
Medium,” Proceedings of the Royal Society A: Mathematical, Physical and Engineering
Sciences, vol. 83, no. 563, pp. 357–365, 1910.
[19] R. A. Millikan, “The isolation of an ion, a precision measurement of its charge, and the
correction of Stokes’s law,” Physical Review (Series I), vol. 32, no. 4, pp. 349–397, 1911.
[20] C. N. Davies, “Definitive equations for the fluid resistance of spheres,” Proceedings of the
Physical Society, vol. 57, no. 4, pp. 259–270, 1945.
[21] ANSYS Inc. (US), “ANSYS Fluent Customization Manual,” vol. 15317, no. January,
pp. 724–746, 2015.
[22] E. K. Vakkilainen, Kraft recovery boilers - principles and practice. Helsinki, Suomen
Soodakattilayhdistys, 2005.
[23] E. K. Vakkilainen, Chemical pulping part 2: Recovery of chemicals and energy, vol. 2.
Helsinki Paper Engineers’ Association/Paperi ja Puu Oy; 2008. p. 28-33, 2008.
[24] ANSYS Inc. (US), “ANSYS FLUENT User ’ s Guide,” p. 2498, 2016.
[25] M. M. Zdravkovich, Flow around circular cylinders: Volume 2: Applications, vol. 2. Oxford
university press, 2003.
[26] Z. Zhan, A. Fry, and J. Wendt, “Relationship between submicron ash aerosol characteristics
and ash deposit compositions and formation rates during air- and oxy-coal combustion,”
Fuel, 2016.
[27] M. C. van Beek, Gas-side fouling in heat-recovery boilers. PhD thesis, Technische Univer-
siteit Eindhoven, 2001.
[28] A. G. Konstandopoulos, “Particle sticking/rebound criteria at oblique impact,” Journal of
Aerosol Science, vol. 37, no. 3, pp. 292–305, 2006.
[29] Z.-X. Tong, M.-J. Li, Y.-L. He, and H.-Z. Tan, “Simulation of real time particle deposition
and removal processes on tubes by coupled numerical method,” Applied Energy, 2016.
[30] H. Han, Y. L. He, W. Q. Tao, and Y. S. Li, “A parameter study of tube bundle heat
exchangers for fouling rate reduction,” International Journal of Heat and Mass Transfer,
vol. 72, pp. 210–221, 2014.
[31] MEMS & Nanotechnology, “Material properties database: Silicon Dioxide (SiO2), film.”
[32] K. L. Johnson, K. Kendall, and A. D. Roberts, “Surface energy and the contact of elas-
tic solids,” in Proceedings of the Royal Society of London A: Mathematical, Physical and
Engineering Sciences, vol. 324, pp. 301–313, The Royal Society, 1971.
[33] P. Mikkanen, Fly ash particle formation in kraft recovery boilers. PhD thesis, VTT Publi-
cations, 2000.

14

View publication stats

You might also like