Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Aerosol Science 129 (2019) 16–27

Contents lists available at ScienceDirect

Journal of Aerosol Science


journal homepage: www.elsevier.com/locate/jaerosci

Experimental and theoretical studies of the relationship between


T
dry and humid normal restitution coefficients

Xue Li, Ming Dong , Sufen Li, Yan Shang
School of Energy and Power Engineering, Key Laboratory of Ocean Energy Utilization and Energy Conservation of Ministry of Education, Dalian
University of Technology, No. 2 Linggong Road, Ganjingzi District, Dalian, Liaoning 116024, China

A R T IC LE I N F O ABS TRA CT

Keywords: The normal restitution coefficient is one of the important parameters in the process of particle
Normal restitution coefficient collision, which is widely used in discrete element modelling (DEM) of gas-solid multiphase
Humid conditions systems. In order to get more accurate results of the simulation, more accurate parameters for
Stokes number DEM modelling are needed. The determination of the normal restitution coefficient for the dry
Physical property
condition is relatively straightforward; however, the unpredictable behaviour of this coefficient
under humid conditions makes its determination more complex. This paper presents an experi-
mental system to determine the normal restitution coefficient using a high-speed camera.
Experiments were performed to measure the incident and rebound velocities of three kinds of fly
ash particles with a diameter of approximately 7 µm. Furthermore, the Stokes number is the key
parameter to calculate the relationship between dry and humid normal restitution coefficients.
The existing theory is suitable for analysing the collision with a thin liquid film; however, the
method is not applicable to the humid conditions. Therefore, a modified Stokes number is pro-
posed to calculate the normal restitution coefficient under these conditions. The normal resti-
tution coefficient is calculated by the modified formula, which is in a good agreement with the
experimental results.

1. Introduction

Particles are widely used in industrial and natural processes including conventional electrostatic precipitators (ESPs)
(Kherbouche, Benmimoun, Tilmatine, Zouaghi, & Zouzou, 2016), circulating fluidized beds (Cai et al., 2016), filtration (Givehchi &
Tan, 2015), and appliance manufacturing industry. Especially the collision-sticking phenomenon among micron-sized particles has
been researched widely. The effects of thermophoresis (Wei et al., 2018), impact angle and particle size (Chen, Li, & Yang, 2015), the
material properties, interfacial properties and structure of the agglomerate (Loreti & Wu, 2018) on particle-wall collision has been
considered by discrete element method (DEM). Liu, Chen, and Li (2018) proposed a modified fluid adhesion number to simulate the
motion of monodisperse spherical microparticles in fluid via a two-way coupled CFD-DEM approach. A liquid bridge is formed in the
humid environment, and the thin liquid film between the two-colliding surfaces play a key role in wet electrostatic precipitator
(WESP) (Dey & Venkataraman, 2012). An approximate theory for the relationship between the dry and wet restitution coefficient
have been proposed by Joseph, Zenit, Hunt, and Rosenwinkel (2001), Davis, Rager, and Good (2002) and Kantak, Galvin, Wildemuth,
and Davis (2005). The Stokes number is an essential parameter in the approximate theory, which characterizes the particle inertia
relative to viscous forces.


Corresponding author.
E-mail address: dongming@dlut.edu.cn (M. Dong).

https://doi.org/10.1016/j.jaerosci.2018.12.006
Received 16 July 2018; Received in revised form 29 November 2018; Accepted 4 December 2018
Available online 07 December 2018
0021-8502/ © 2018 Elsevier Ltd. All rights reserved.
X. Li et al. Journal of Aerosol Science 129 (2019) 16–27

Recently, there have been some experimental and numerical studies on particle collision with the thin liquid film or under humid
condition. Gollwitzer, Rehberg, Kruelle, and Huang (2012) conducted experimental studies of the restitution coefficient on the
impact velocity and various properties of the particle and liquid. The viscosity of the thin liquid film had a negligible effect on the
restitution coefficient. The influence of the liquid viscosity and layer thickness on the restitution coefficient were analysed by Crüger,
Heinrich, Antonyuk, Deen, and Kuipers (2016) through the improved experimental approach with two synchronized high-speed
cameras capturing the collision behaviour of a particle in three dimensions. Dörmann and Schmid (2017), Landi, Barletta, and Poletto
(2011) and Mu and Su (2007) analysed the energy loss, liquid bridge force and liquid bridge geometry caused by the humid condition
during particle collisions. Due to the hindering effect of the liquid bridge, the energy loss and liquid bridge force of particle increased
during a collision. Givehchi and Tan (2015), Forsyth, Hutton, and Rhodes (2002) and Dörmann and Schmid (2014) investigated the
behaviour of liquid bridge force and elastic-plastic impaction in a collision under humid condition through the experiment and
theoretical analysis. A direct numerical simulation model was used to simulate solid–liquid–gas three-phase flow by Washino, Tan,
Hounslow, and Salman (2013), and they calculated the liquid bridge force on particle under humid condition. Tang, He, Tai, and Wen
(2017) investigated the particle flow behaviour in dry and humid multiple-spout fluidized beds through the experiment and simu-
lation method. The research results showed that the energy loss of particle and the effect of drag force became larger under humid
conditions. Zhang and Li (2017) combined discrete element model and computational fluid dynamics to analyse wet particle flows in
a thin spouted bed by experimental and modelling approaches.
The Stokes number is an important parameter, which is used to determine the wet restitution coefficient based on the dry
restitution coefficient. There is a connection between dry restitution coefficient and wet restitution coefficient. Davis et al. (2002) and
Davis, Serayssol, and Hinch (1986) set up an experimental system to measure the incident and rebound velocities of small plastic and
metal spheres impacting the surface covered with a thin liquid film. The wet restitution coefficient calculated by Stokes number,
elasticity parameter, and dry restitution coefficient show good agreement with the experimental results. Adams, Thornton, and Lian
(1996) estimated the wet restitution coefficient based on head-on elastohydrodynamic collisions approaching the wetted target
surface, coupled with Hertzian elasticity theory. The Stokes number had been corrected to suit for different situations. Gollwitzer
et al. (2012) proposed a modified Stokes number to revise the liquid viscosity of the thin liquid film, and Ma, Liu, and Chen (2013,
2016) presented another modified Stokes number to correct both the liquid viscosity and layer thickness of the thin liquid film.
In literature, the Stokes number has been modified for the collision with thin liquid film; however, a model for the particle
collision under humid condition has never been proposed before. In this paper, an experimental method and modified formulas are
presented to determine the laws of dry restitution coefficient and wet restitution coefficient under different humid conditions. In
Section 2, the experimental set-up for measuring the incident and rebound velocities of particles and analysing the material prop-
erties of fly ash particles is presented. In Section 3, the modified Stokes number and elasticity parameter is introduced to calculate the
wet restitution coefficient under different humid conditions. In Section 4, the effects of physical properties and the humidity on the
particle collision are experimentally determined. Then, the wet restitution coefficient is calculated by the modified formula, which is
in a good agreement with the experimental results.

2. Experimental analysis

2.1. Experimental system

In the present study, the experimental facility developed by Dong, Li, Mei, and Li (2017), was used to investigate the impact of
micro-particles on planar surfaces under dry and humid conditions. The system is used for measuring the incident and rebound
velocities of particles as shown in Fig. 1.
The experimental system is composed of four parts: inlet system, humidity control system, collision unit, and high-speed camera
system. In this experiment, the particles were injected using a particle generator. The particles entered the collision regions by the
blend of dry and humidity airflows. BGI device (atomizer aerosol generator, BGI Inc., USA) is used to provide the moisture en-
vironment. The dry-humidity mixed flow can provide the required humid condition; therefore, the humidity and velocity could be
changed by adjusting the mass flow meters, which can control the mass flow of fluid. The humidity of the outlet gases is measured

Fig. 1. Schematic of the experimental configuration.

17
X. Li et al. Journal of Aerosol Science 129 (2019) 16–27

Fig. 2. Surface roughness using FlatMaster200.

using a hygrometer. The fly ash particle vertically impacted onto the planar surface which is made of stainless-steel cylinder with a
diameter of 2 mm. The planar surface with 9.2 nm RMS roughness was observed by FlatMaster200 (Corning Tropel, USA), as shown
in Fig. 2. A high-speed camera system (Phantom V12.1, AMETEK Company, USA) was used to determine the incident and rebound
velocities of each particle, and particle diameter. The images were shown in Fig. 3, the size and incident velocity of the particle in the
picture is 7 µm and 5.3 m/s, respectively.
The liquid bridge will be formed between the particle and the planar surface during a collision in Fig. 4. The equation of liquid
bridge volume (V) and separation distance (D*) are as follows (Willett et al., 2000):

V = πa (4r 2c 2 − D 2) (1)

θ + θ2 ⎞ 1/3 2 2/3
D* = ⎛1 + 1 ⎛V − V ⎞
⎝ 8 ⎠⎝ 5 ⎠ (2)

cos θ1 + cos θ2
c=
2 (3)

where, a is particle radius; r, θ1, and θ2 are the radii of curvature of the meniscus and the contact angles which could be recognized in
Fig. 4. At constant vapor pressure and in equilibrium the radius is constant and can easily be calculated with Kelvin's equation (Butt &
Kappl, 2009).

λK
r=−
ln(P / P0) (4)

P/P0 is the relative vapor pressure. The values of the Kelvin length λK (Butt, 2008), θ1 and θ2 are 0.523 nm, 20° and 40° (Bateman,
Belassein, & Martin, 2014), respectively. According to equations given above, the separation distances of fly ash are 18 nm, 21 nm,
and 24.7 nm for LN, XJ and NM coals, respectively. The separation distance is larger than the surface roughness, so surface roughness
could be covered by liquid bridge. Furthermore, the liquid can increase lubricity and reduce friction, which could reduce the effect of
surface roughness.
The particles were recorded using the high-speed digital camera having a resolution of 256 × 128 at an exposure time of 6.22 μs.
The frame rate was 120,000 frames/s which could capture the micro-particle trajectories. The particle velocity and experimental
humidity were obtained by adjusting the mass flow meters. The range of particle incident velocity was 1–7 m/s. The experimental
humidity was dry, 35%, 50%, and 65%. All experiments were conducted at a room temperature of 20 ± 2 °C. For avoiding particle
aggregation, the particles were stored in a drying oven at 110 °C for 5 h before an experiment. The experimental data was selected the
single particle to measure the incident and rebound velocity. Under proper filter data, the difference between diameter of micro-
capsules is less than 5%. The effect of the gas flow on the collision process, experimental procedure and velocity measurement
method was discussed in detail in the literature (Dong et al., 2017), so the present paper will not describe this.

Fig. 3. Typical recorded image for the impact of a fly ash on a flat surface.

18
X. Li et al. Journal of Aerosol Science 129 (2019) 16–27

Fig. 4. The liquid bridge between particle and planar surface.

2.2. Particles

In the experiment, bituminous coal from Fushun, Liaoning Province (LN), meager coal from Zhundong, Xinjiang Province (XJ),
and anthracite coal from XinlinGol, Neimeng Province (NM) were used to investigate the effects of physical properties. The types of
coals were selected by analysing the volatile matter content of samples (Wu, 2014). The coals are turned into fly ash particles using a
muffle furnace. The total elemental analysis of different fly ash particles was performed from the X-ray fluorescence spectrum, shown
in Table 1.
The content of three fly ash particles is quite different; therefore, the material properties are different as well. The particle size
distribution range of fly ash particles was comparatively centralized. The bulk of the diameters determined by a laser scattering
particle analyser ranged from 1 to 15 µm, as shown in Fig. 5. The particle diameter could be measured by the Phantom software PCC;
therefore, the different size particles could be selected. The particle diameter selected in the experiment analysis are 7 µm, which are
much larger than the surface roughness. Therefore, the surface roughness could be ignored (Wall, John, Wang, & Goren, 1990).
The components of the samples were analysed by X-ray diffraction; in addition, the content of the crystallization phase and
amorphous phase were measured in Fig. 6, and the data are presented in Table 2. MDI Jade software was used to calculate the weight
percentage of fly ash crystallinity.
The physical parameters were calculated using the following equations, and the results are listed in Table 3 (Matsunaga, Kim,
Hardcastle, & Rohatgi, 2002).

ρflyash = ρSiO2 VSiO2 + ρCaSO4 VCaSO4 + ρgalss Vglass (5)

Eflyash = ESiO2 VSiO2 + ECaSO4 VCaSO4 + Egalss Vglass (6)

The density and Young's modulus values are smallest for the LN fly ash. It shows that the inertia force and the ability to resist
elastic deformation of the fly ash for LN particle are small, the ability to keep the original movement tendency and resist deformation
is small as well during the collision of a particle. There is a small range of Young's modulus value of XJ and NM fly ashes, but the
range of density is large.

3. Theoretical analysis

The main energy loss during particle collision under dry conditions consists of plastic deformation (Johnson, 1986), elastic waves
(Hunter, 1957), viscoelasticity of the solids (Falcon et al., 1998; Ramírez et al., 1999), and adhesive forces (Dahneke, 1971). Under
wet conditions, the energy loss is further increased by viscous losses in the thin fluid film between the colliding surfaces. Further-
more, the wet restitution coefficient could be obtained from the dry restitution coefficient. Davis et al. (2002) gave the expression of a
closed form for the wet restitution coefficient based on an approximate theory:

Table 1
The total elemental analysis of fly ash particles.
Content% SiO2 Al2O3 CaO MgO TiO2 SO3 P2O5 Na2O Fe2O3 Others

LN 52.30 36.40 2.39 0.63 2.84 2.08 1.42 0.56 0 1.38


XJ 28.20 12.30 26.50 1.82 0.30 9.51 0.01 2.48 9.10 9.78
NM 14.30 15.70 23.40 9.00 0.60 17.20 0.13 4.82 12.4 2.45

19
X. Li et al. Journal of Aerosol Science 129 (2019) 16–27

Fig. 5. Distribution of particle size.

Fig. 6. Fly ash X-ray diffraction patterns.

Table 2
Crystalline volume fractions.
Fly ash Crystalline phase (vol%) Amorphous phase (vol%)

SiO2 CaSO4 Al2O3

LN 18.43 – 12.57 69
XJ 2.24 66.53 – 31.23
NM 28.27 61 – 10.73

Table 3
Physical parameters of fly ash particles.
Volume fraction

Fly ash SiO2 (%) CaSO4 (%) Glass phase (%) elastic yield limit (GPa) Density (kg/m3) Young's modulus (GPa)

LN 18.43 12.57 69 3.81 2680 116.72


XJ 2.35 62.76 34.88 3.76 2952.69 139.39
NM 28.93 58.99 12.08 3.77 2814.74 141.01

(
e wet = edry 1 −
StC
St ) for St > StC
e wet = 0 for St ≤ StC (7)

where, ewet and edry are the wet and dry restitution coefficients, respectively. StC is the critical Stokes number above which the
1
rebound occurs. Based on the Eq. (3), a ‘master curve’ for ewet/edry versus St/StC agrees with the curve y = 1 − x (Davis et al., 2002).
And many scholars calculated the wet restitution coefficients by this method with diameter ranged from 0.05 mm to 10 mm (Gondret,
Hallouin, Lance, & Petit, 1999; Gondret, Lance, & Petit, 2002; Joseph et al., 2001). The Eq. (7) is mainly corrected by Stokes number
that it is a reasonable modification for the solid aerosol particles with a few microns (Rh, Jm, & Ej, 2006). Furthermore, the density of
the three kinds of particle is 2680 kg/m3, 2952.69 kg/m3 and 2814.74 kg/m3 respectively which are much larger than the density of

20
X. Li et al. Journal of Aerosol Science 129 (2019) 16–27

water. Therefore, it is suitable for the particle size in this paper.


The Stokes number, defined as the ratio between the inertia of the particle and the viscosity of the liquid, is used to characterize
the dependence of the restitution coefficient on various control parameters. The expression has been defined by many researchers for
different conditions. Davis et al. (2002) proposed a modified Stokes number:
mv0
StD =
6πμa2 (8)
where m, v0, μ and a are particle mass, the incident velocity of the particle, the viscosity of thin fluid layer and particle radius,
respectively. The Eq. (8) applies to the collision with a thin fluid layer. The energy loss is caused by the formation of the liquid bridge
between particle and plate surface during a collision. It is affected by the thickness and viscosity of the fluid layer. In order to consider
the layer thickness, Ma et al. (2013) proposed a modified Stokes number expression:
mv0
StN =
μdh (9)
where h is the thickness of the fluid layer. The liquid viscosity and thickness conditions are considered in this equation.
As environmental humidity increases, the liquid bridge is formed on the contact surface by capillary condensation. Therefore, the
humidity influences the particle collision process. However, the corrected formula of the Stokes number is not appropriate for
characterizing the rebound behaviour under humid condition.
Therefore, a modified Stokes number is proposed to consider the influence of environmental humidity. The Stokes number can be
presented as follows:
mv0
StRH =
6πμa2e RH (10)
RH
where RH is the environmental humidity; e represent the humidity correction factor; m is the particle mass; v0 is the incident
velocity; a is the particle radius; μ is the liquid viscosity. When the humidity exists, the liquid bridge causes the viscous force.
To compute the critical Stokes number in Eq. (7), the elasticity parameter ε need to be calculated. The elasticity parameter
describes the deformation of the particle in the elastic process (Davis et al., 1986). The elasticity parameter is a ratio of viscous forces,
that causes deformation, and the stiffness of the solids, that resists deformation. A higher value of ε indicates an easier deformation of
the particle.
3Kμv0 a3/2
ε=
πx 05/2 (11)
Where x0 is the initial separation between the undeformed spheres. K is defined as follows:

4 ⎛ (1 − υ12) (1 − υ22) ⎞
K= ⎜ + ⎟
3 ⎝ E1 E2 ⎠ (12)
where Ei and νi are Young's modulus and Poisson's ratio, respectively. When a particle is impacting the surface with thin liquid film, x0
is defined as (Davis et al., 1986):
x 0 = 0.01a (13)
where two modified methods were presented to calculated the elasticity parameter.
x 0 = 0.01aRH (14)

x 0 = 0.01a (e RH − 1) (15)
By comparison the value of the calculations and experiments, the optimum expression could be chosen.
Because the elasticity parameters are calculated in the elastic stage, the particle velocity should be less than the limiting elastic
velocity. The elastic yield limit is only attained for incident velocities greater than a limiting elastic velocity, Vy, given by (Johnson,
Kendall, & Roberts, 1971):
1
2π 2 2 2 5
Vy = ⎛ ⎞ ⎜⎛ ⎟⎞ Y 2
⎝ 3K ⎠ ⎝ 5ρ ⎠ (16)
where Y is the elastic yield limit. According to the value of parameters given in Table 3, the limiting elastic velocities of fly ash are
4.189 m/s, 4.072 m/s and 4.126 m/s for LN, XJ and NM coals, respectively.
The compliance parameter χ is used to calculate the critical Stokes number StC which defined as the ratio between the elasticity
parameter and the Stokes number.
ε 18πKμ2 a3.5e RH
χ= =
St mx 02.5 (17)
The compliance parameter depends only on material properties. When the minimum value of the Stokes number is close to 2, the
StC can be calculated as follows (Kantak, 2005).

21
X. Li et al. Journal of Aerosol Science 129 (2019) 16–27

StC = −1.2 − 0.4 ln(χ ) (18)


Using the modified Stokes number and the compliance parameter, the restitution coefficient under humid conditions could be
calculated.
The main adhesion force of the fine particle is the van der Waals force. The force would be reduced by the liquid bridge under
humidity condition. The equation of the van der Waals force is defined as (Butt, Farshchitabrizi, & Kappl, 2006):

(AHl − AHg )(d + 1 − 2 cos β ) A l (d − 1) AHg


Fvdw = − 2
+ H 2

6a (d + 1 − cos β ) 6ad 6a (d + 1) (19)

D
d=
a (20)

l2
cos β = 1−
a2 (21)

2
λK ⎛ λK ⎞ − FC λK
l= + ⎜ ⎟
2 ln(RH ) ⎝ 2 ln(RH ) ⎠ πγ ln(RH ) (22)
where AHl and AHg are the Hamek constant among the contacting surfaces. L, r, D, and γ are the radii of curvature of the meniscus, the
separation distance and water surface tension coefficient which could be recognized in Fig. 4. FC is the capillary force, given by (Butt
& Kappl, 2009):

D
FC = 2πγa ⎜⎛2c − ⎞⎟
⎝ r⎠ (23)
The values of van der Waals force and capillary force versus the separation distance are shown in Fig. 7. The magnitude of van der
Waals force and capillary force is 10−7 and 10−6, respectively under 35% humid condition. So the effect of surface adhesion could be
ignored. In this paper, we only consider the viscous force caused by the humidity condition.

4. Results and discussion

4.1. Experimental results

The normal restitution coefficient en is calculated as a ratio of the normal rebound velocity to the normal incident velocity
determined experimentally. The experiment was performed five times at the same conditions. Several experimental data have been
obtained at the same conditions, and forming the error curve.
Fig. 8 shows the normal restitution coefficient versus the incident velocity for three kinds of fly ashes under different humid
conditions. The normal restitution coefficient depends on the humid condition, which is enhancing the particle deposition. The
normal restitution coefficient decreases with humidity increasing due to the viscous losses.
From the normal restitution coefficient in Fig. 8, the value of fly ash for LN particle has the maximum at all conditions, and that is
because the density and Young's modulus for LN particles are smaller than another two kinds of particles. The inertial force increases
with density increasing, and the energy loss will increase accordingly during a collision. Young's modulus reflects the ability to resist
elastic deformation; thus, the elastic deformation decreases with Young's modulus increasing. So, the LN fly ash has less energy loss
and larger storage of elastic energy, therefore, the normal restitution coefficient is greater. The Young's modulus of XJ coal particles is

Fig. 7. Van der Waals force and capillary force versus the separation distance under 35% humid condition.

22
X. Li et al. Journal of Aerosol Science 129 (2019) 16–27

Fig. 8. Normal restitution coefficient versus normal incident velocity component for LN, XJ, and NM fly ash particles under different humid
conditions.

close to that of NM coal particles, therefore, the normal restitution coefficient has the minimum for the XJ particles, because of the
larger density.

4.2. Stokes number

The Stokes number St represents the characteristic of the particle in the flow field, which is defined as the inertia of the particles
relative to the viscous forces. A modified Stokes number is proposed to describe the particle motion under humid conditions.
The normal restitution coefficient versus the Stokes number is shown in Fig. 9. The normal restitution coefficient first increases
and then decreases with Stokes number increasing. As the humidity increases, the distribution of normal restitution coefficient
becomes narrow, but the maximum value is not changed. To achieve the same Stokes number, the inertia and incident velocity of the
particles should increase as the humidity, and the maximum incident velocity should remain constant at different conditions, so the
curve becomes narrow. The difference among the normal restitution coefficients is little with the Stokes number in the range of 2–4.
While the difference among the normal restitution coefficients is large with Stokes number increasing. This phenomenon is caused by
the changes of the elasticity parameter.
Fig. 10 shows the modified Stokes number versus the incident velocity. The orange vertical line represents the limiting elastic
velocities of the particle. The corresponding Stokes number is the point at which the normal restitution coefficient begins to decrease
in Fig. 9. The difference among the normal restitution coefficients is small in the elastic stage, but it becomes larger in the plastic
stage. Thus, the motion characteristic and energy loss of the fly ash particles are in accordance with the same Stokes number in the
elastic stage. The energy loss caused by the plastic deformation increases with incident velocity in the plastic stage increasing. The
viscous force increases with humidity increasing. To achieve the same Stokes number, the inertia force and incident velocity should
be larger. Therefore, as the humidity increases, the normal restitution coefficient decreases faster at the same Stokes number.

4.3. Elasticity parameter

The elasticity parameter ε is used to describe the deformation of the particle in a collision. The elasticity parameter versus the
incident velocity calculated by Eq. (11) is shown in Fig. 11. In the elastic stage, the elasticity parameter increases with incident
velocity of the particle increasing, which means the deformation of the particle is increased in a collision. However, as the humidity
increases, the effect of incident velocity on the elasticity parameter is reduced. The effect of humidity is relatively more significant.
With the humidity increasing, the viscous force caused by the liquid bridge has greatly increased under the condition of relative lower
humidity.
From Fig. 11, the value of elasticity parameter for XJ particles is close to the value for NM particles under the same humid
condition which means the deformation for XJ particles is similar to that for NM particles. Furthermore, the stored elastic potential

Fig. 9. The normal restitution coefficient versus the Stokes number.

23
X. Li et al. Journal of Aerosol Science 129 (2019) 16–27

Fig. 10. Modified Stokes number versus the incident velocity.

Fig. 11. The elasticity parameter versus the incident velocity.

energy is closer at the same incident velocity. Thus, the particles have the same stored energy and rebound velocity. So, the normal
restitution coefficient values of fly ash for XJ and NM particles are closer under the humid condition which could be seen in Fig. 8.

4.4. The wet restitution coefficients calculation

The wet and dry restitution coefficients are shown in Eq. (7). The Stokes number is affected by the humidity level. The critical
Stokes number could be obtained from Eq. (18). Two modified forms have been exhibited in Eqs. (14) and (15). Because the elasticity
parameters are applicable for the elastic stage, the velocity below the limiting elastic velocity is calculated only. The modified
compliance parameters are as follows.
ε 18πKμ2 a3.5e RH
χ= =
St m (0.01aRH )2.5 (24)

ε 18πμ2 a3.5e RH
χ= =
St m [0.01a (e RH − 1)]2.5 (25)
The comparison of the two modified forms is shown in Fig. 12.
The wet restitution coefficients calculated by Eq. (25) are closer to the experimental results. The force and deformation of the
particle are nonlinear functions of the humidity. Because of the variation of the elasticity parameter, the humidity effect on the
particle deformation decreases at higher humid conditions.
The intersections between the straight line and the horizontal x-axis are the critical Stokes numbers, as shown in Fig. 13. The

Fig. 12. The comparison of the two modified forms with the experimental data.

24
X. Li et al. Journal of Aerosol Science 129 (2019) 16–27

Fig. 13. The experimental values of critical Stokes numbers for LN, XJ and NM fly ash particles.

critical Stokes numbers for LN and NM particles are close to 2, however, the value of XJ is much higher than 2. If Eq. (18) is used to
calculate the critical Stokes number, a large error can be expected. The calculation of critical Stokes numbers for XJ particles is
illustrated in Eq. (24) (Davis et al., 2002).
StC = −0.2 − 0.4 ln(χ ) − 0.4 ln(StC ) (26)

Using Eq. (26), the critical Stokes number could be calculated. The comparison of experimental and calculated results is shown in
Fig. 14.
The results indicate that the experimental and calculated values are close. The maximum error was 10% at 65% humidity
condition for XJ particles, while the average error is 4.83%. Because the critical Stokes number calculated by Eq. (18), the error is
relatively large. The difference between the experimental and calculated data is within the experimental error range. When the
incident velocity is small, the velocity is close to the critical capture velocity, the normal restitution coefficient changes faster with
the incident velocity increasing, therefore, the experiment error exists. The curve of equivalent normal restitution coefficient and
1
equivalent Stokes number follow the function y = 1 − x (Davis et al., 2002), which is shown in Fig. 15. So, it is reasonable to
calculate the wet restitution coefficients with the modified formula.

5. Conclusions

In this study, a detailed model for investigation of particle-wall collision mechanics under the humid condition is proposed and
compared with the experimental results. The modified Stokes number is proposed to represent the relationship between dry and
humid normal restitution coefficients. The following conclusions can be drawn from this study.
First, the normal restitution coefficients of three kinds of fly ashes from Liaoning Province (LN), Xinjiang Province (XJ) and
Neimeng Province (NM) under dry, 35%, 50%, and 60% humidity conditions were obtained experimentally. Results showed that the
normal restitution coefficient decreases with relative humidity, particle density, and Young's modulus increasing.
Second, the normal restitution coefficient versus the modified Stokes number was obtained. The difference among the normal
restitution coefficients is small for the Stokes numbers in the range of 2–4. While the difference among the normal restitution
coefficients is larger for increasing the Stokes numbers.
Third, the deformation of the particle decreases with the humidity increasing due to the hindering effect of the liquid bridge. With
the humidity increasing, the effects of incident velocity and humidity on the elasticity parameter are reduced. The value of elasticity
parameter for XJ particles is closer to the value for NM particle, therefore, their values of normal restitution coefficient are similar

Fig. 14. The comparison of experimental and calculated results.

25
X. Li et al. Journal of Aerosol Science 129 (2019) 16–27

Fig. 15. Equivalent normal restitution coefficient versus equivalent Stokes number.

under humid conditions.


Finally, the wet restitution coefficient calculated by the modified formula and the experimental data are close. The curve of
equivalent normal coefficient restitution versus equivalent Stokes number agrees with the function proposed by Davis et al. (2002).

Acknowledgments

The authors acknowledge the National Natural Science Foundation of China (No. 51576030) and the National Key Research and
Development Program of China (No. 2016YFB0600602). The authors give special thanks to Professor Kazimierz Adamiak of the
Department of Electrical and Computer Engineering in Western University for his help with the English language.

References

Adams, M. J., Thornton, C., & Lian, G. (1996). Elastohydrodynamic collisions of solid spheres. Journal of Fluid Mechanics, 311, 141–152.
Bateman, A. P., Belassein, H., & Martin, S. T. (2014). Impactor apparatus for the study of particle rebound: Relative humidity and capillary forces. Aerosol Science and
Technology, 48, 42–52.
Butt, H. (2008). Capillary forces: Influence of roughness and heterogeneity. Langmuir the Acs Journal of Surfaces & Colloids, 24, 4715–4721.
Butt, H., Farshchitabrizi, M., & Kappl, M. (2006). Using capillary forces to determine the geometry of nanocontacts. Journal of Applied Physics, 100, 359–392.
Butt, H. J., & Kappl, M. (2009). Normal capillary forces. Advances in Colloid & Interface Science, 146, 48–60.
Cai, J., Wu, C., Moghtaderi, B., Doroodchi, E., Peng, Z., Zhao, X., & Yuan, Z. L. (2016). A numerical study of the orientation of cylindrical particles in a circulating
fluidized bed. Industrial & Engineering Chemistry Research, 55, 12806–12817.
Chen, S., Li, S., & Yang, M. (2015). Sticking/rebound criterion for collisions of small adhesive particles: Effects of impact parameter and particle size. Powder
Technology, 274, 431–440.
Crüger, B., Heinrich, S., Antonyuk, S., Deen, N. G., & Kuipers, J. A. M. (2016). Experimental study of oblique impact of particles on wet surfaces. Chemical Engineering
Research & Design, 110, 209–219.
Dahneke, B. (1971). The capture of aerosol particles by surfaces. Journal of Colloid & Interface Science, 37, 342–353.
Davis, R. H., Rager, D. A., & Good, B. T. (2002). Elastohydrodynamic rebound of spheres from coated surfaces. Journal of Fluid Mechanics, 468, 107–119.
Davis, R. H., Serayssol, J. M., & Hinch, E. J. (1986). The elastohydrodynamic collision of two spheres. Journal of Fluid Mechanics, 163, 479–497.
Dey, L., & Venkataraman, C. (2012). A wet electrostatic precipitator (WESP) for soft nanoparticle collection. Aerosol Science & Technology, 46, 750–759.
Dong, M., Li, X., Mei, Y., & Li, S. (2017). Experimental and theoretical analyses on the effect of physical properties and humidity of fly ash impacting on a flat surface.
Journal of Aerosol Science, 117, 85–99.
Dörmann, M., & Schmid, H. J. (2014). Simulation of capillary bridges between particles. Langmuir the Acs Journal of Surfaces & Colloids, 30, 1055–1062.
Dörmann, M., & Schmid, H. J. (2017). Distance-dependency of capillary bridges in thermodynamic equilibrium. Powder Technology, 312, 175–183.
Falcon, E., Laroche, C., Fauve, S., & Coste, C. (1998). Behavior of one inelastic ball bouncing repeatedly off the ground. The European Physical Journal B - Condensed
Matter and Complex Systems, 3, 45–57.
Forsyth, A. J., Hutton, S., & Rhodes, M. J. (2002). Effect of cohesive interparticle force on the flow characteristics of granular material. Powder Technology, 126,
150–154.
Givehchi, R., & Tan, Z. (2015). The effect of capillary force on airborne nanoparticle filtration. Journal of Aerosol Science, 83, 12–24.
Gollwitzer, F., Rehberg, I., Kruelle, C. A., & Huang, K. (2012). Coefficient of restitution for wet particles. Physical Review E Statistical Nonlinear & Soft Matter Physics, 86,
1193–1222.
Gondret, P., Hallouin, E., Lance, M., & Petit, L. (1999). Experiments on the motion of a solid sphere toward a wall: From viscous dissipation to elastohydrodynamic
bouncing. Physics of Fluids, 11, 2803–2805.
Gondret, P., Lance, M., & Petit, L. (2002). Bouncing motion of spherical particles in fluids. Physics of Fluids, 14, 643–652.
Hunter, S. C. (1957). Energy absorbed by elastic waves during impact. Journal of the Mechanics & Physics of Solids, 5, 162–171.
Johnson, K. L. (1986). Contact mechanics. Journal of Tribology, 108, 464.
Johnson, K. L., Kendall, K., & Roberts, A. D. (1971). Surface energy and the contact of elastic solids. Proceedings of the Royal Society of London A Mathematical Physical &
Engineering Sciences, 324, 301–313.
Joseph, G. G., Zenit, R., Hunt, M. L., & Rosenwinkel, A. M. (2001). Particle-wall collisions in a viscous fluid. Journal of Fluid Mechanics, 433, 329–346.
Kantak, A. A. (2005). Wet particle collisions. University of Colorado: ProQuest Information and Learning Company.
Kantak, A. A., Galvin, J. E., Wildemuth, D. J., & Davis, R. H. (2005). Low-velocity collisions of particles with a dry or wet wall. Microgravity – Science and Technology,
17, 18–25.
Kherbouche, F., Benmimoun, Y., Tilmatine, A., Zouaghi, A., & Zouzou, N. (2016). Study of a new electrostatic precipitator with asymmetrical wire-to-cylinder

26
X. Li et al. Journal of Aerosol Science 129 (2019) 16–27

configuration for cement particles collection. Journal of Electrostatics, 83, 7–15.


Landi, G., Barletta, D., & Poletto, M. (2011). Modelling and experiments on the effect of air humidity on the flow properties of glass powders. Powder Technology, 207,
437–443.
Liu, W., Chen, S., & Li, S. (2018). Random adhesive loose packings of micron-sized particles under a uniform flow field. Powder Technology, 335, 70–76.
Loreti, S., & Wu, C. Y. (2018). Three-dimensional discrete element modelling of three point bending tests: The effect of surface energy on the tensile strength. Powder
Technology, 10, 1–8.
Ma, J., Liu, D., & Chen, X. (2013). Experimental study of oblique impact between dry spheres and liquid layers. Physical Review E Statistical Nonlinear & Soft Matter
Physics, 88, 033018.
Ma, J., Liu, D., & Chen, X. (2016). Normal and oblique impacts between smooth spheres and liquid layers: Liquid bridge and restitution coefficient. Powder Technology,
301, 747–759.
Matsunaga, T., Kim, J. K., Hardcastle, S., & Rohatgi, P. K. (2002). Crystallinity and selected properties of fly ash particles. Materials Science & Engineering A Structural
Materials Properties Microstructure & Processing, 325, 333–343.
Mu, F., & Su, X. (2007). Analysis of liquid bridge between spherical particles. Particuology, 5, 420–424.
Ramírez, R., Pöschel, T., Brilliantov, N. V., & Schwager, T. (1999). Coefficient of restitution of colliding viscoelastic spheres. Physical Review E Statistical Physics Plasmas
Fluids & Related Interdisciplinary Topics, 60, 4465.
Rh, D., Jm, S., & Ej, H. (2006). The elastohydrodynamic collision of two spheres. Journal of Fluid Mechanics, 163, 479–497.
Tang, T., He, Y., Tai, T., & Wen, D. (2017). DEM numerical investigation of wet particle flow behaviors in multiple-spout fluidized beds. Chemical Engineering Science,
172, 79–99.
Wall, S., John, W., Wang, H.-C., & Goren, S. L. (1990). Measurements of kinetic energy loss for particles impacting surfaces. Aerosol Science & Technology, 12, 926–946.
Washino, K., Tan, H. S., Hounslow, M. J., & Salman, A. D. (2013). A new capillary force model implemented in micro-scale CFD–DEM coupling for wet granulation.
Chemical Engineering Science, 93, 197–205.
Wei, M. Z., Zhang, Y. Y., Luo, X. W., Li, X. W., Wu, X. X., & Zhang, Z. M. (2018). Graphite dust deposition on HTGR steam generator: Effects of particle-wall and
particle-vortex interactions. Nuclear Engineering and Design, 330, 217–224.
Willett, Christopher D., Adams, Michael J., Johnson, Simon A., & Seville, J. P. K. (2000). Capillary Bridges between Two Spherical Bodies. Langmuir, 16, 9396–9405.
Wu, W. L. (2014). Boiler and boiler-room equipment. China Architecture & Building Press.
Zhang, H., & Li, S. (2017). DEM simulation of wet granular-fluid flows in spouted beds: Numerical studies and experimental verifications. Powder Technology, 318,
337–349.

27

You might also like