Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Case Studies in Thermal Engineering 49 (2023) 103354

Contents lists available at ScienceDirect

Case Studies in Thermal Engineering


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

Optimizing fire extinguishing effects of pneumatic sandblasting


fire extinguishers based on experiments and CFD-DEM
coupling simulations
Fanbao Chen a, b, c, Tingting Xu d, Guanyu Hou a, b, c, Jianhua Huang a, b, c,
Guoqing Zhu a, b, c, *, Tao Deng e, Zhenhua Jiang e, Ziyang Wang e
a
Jiangsu Key Laboratory of Fire Safety in Urban Underground Space, China University of Mining and Technology, Xuzhou, 221116, China
b
Fire Research Institute, China University of Mining and Technology, Xuzhou, 221116, China
c
School of Safety Engineering, China University of Mining and Technology, Xuzhou, 221116, China
d
China Academy of Urban Planning and Design, Beijing, 100044, China
e
Department of Mechanical Engineering, The University of Auckland, Auckland, 1142, New Zealand

A R T I C L E I N F O A B S T R A C T

Keywords: This study focuses on addressing specific issues and research gaps related to pneumatic sand­
CFD-DEM coupling blasting effects for forest fire prevention and control. The primary objective is to determine the
Pneumatic extinguisher sensitivity of geometric parameters and recommend optimal design parameters for the extin­
Sandblasting guisher. The investigation employs a combination of Computational Fluid Dynamics-Discrete
Forest fire Element Methods (CFD-DEM) coupling simulations and experiments. The simulation and exper­
imental results suggest flat outlets have the farthest sandblasting distance. Further experimen­
tation shows that the outlet with different diameters show different sandblasting distances and
accumulation effects, and the increased air duct length leads to decreased average airflow ve­
locity at the outlet. In addition, the blade number of impellers was also studied for a longer
sandblasting distance. Finally, the study found that the optimal parameters (79 cm air duct with
the flat outlet of 8 cm diameter, 20-blade impeller), resulted in longer fire extinguishing length in
fire extinguishing experiments compared to the original extinguisher. The findings provide
valuable insights for improving the effectiveness of fire extinguishers and enhancing forest fire
management strategies. The research outcomes have the potential to make a substantial impact in
mitigating the devastating effects of forest fires and safeguarding natural resources and human
lives.

Nomenclature

→r Particle centroid vector position, m


a Contact radius, m
r∗ Particle equivalent radius, m
Fn Normal force, N

* Corresponding author. Jiangsu Key Laboratory of Fire Safety in Urban Underground Space, China University of Mining and Technology, Xuzhou, 221116, China.
E-mail addresses: cfb119xz@cumt.edu.cn (F. Chen), 912686444@qq.com (T. Xu), 2802725023@qq.com (G. Hou), 13165911690@163.com (J. Huang),
zgq119xz@cumt.edu.cn (G. Zhu), tden033@aucklanduni.ac.nz (T. Deng), zjia709@aucklanduni.ac.nz (Z. Jiang), zwan472@aucklanduni.ac.nz (Z. Wang).

https://doi.org/10.1016/j.csite.2023.103354
Received 29 May 2023; Received in revised form 18 July 2023; Accepted 30 July 2023
Available online 31 July 2023
2214-157X/© 2023 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

E∗ Equivalent elastic modulus, Pa


Fnd Normal damping force, N
m∗ Equivalent mass of particle, kg
vrn Normal component of relative particle velocity, m/s
Sn Normal stiffness, N/m

n Normal unit vector
St Tangential stiffness, N/m
G∗ Equivalent shear modulus, Pa
Ftd Tangential damping force, N
Tt Tangential moment, N/m
Tr Frictional moment, N/m
u
→ Fluid velocity, m/s
mp Mass of particle, kg
up →
→ up Velocity of particle, m/s

F Additional force, N
→u− →
u
mp τ r p Drag force, N
Re Relative Reynolds number
Ip Moment of inertia, m4
Cω Rotational drag coefficient of particle

T Torque applied to particle in fluid domain, N/m
up,y Particle velocity in the y-direction of Cartesian coordinate, m/s
uy Fluid velocity in the y-direction of Cartesian coordinate, m/s
up,x Particle velocity in the x-direction of Cartesian coordinate, m/s
ux Fluid velocity in the x-direction of Cartesian coordinate, m/s

Greek
εn Normal overlap value
μ1 、 μ2 Poisson’s ratio
εt Tangential overlap value
μs Static friction coefficient
ρ Fluid density, kg/m3
ρp Particle density, kg/m3
τr Particle relaxation time, s
μ Fluid viscosity, Pa⋅s
̅→
ωp Particle angular velocity, rad/s

Ω Angular velocity of particle-fluid, rad/s

Subscripts
CFD Computational fluid dynamics
MRF Multiple reference frame
RANS Reynolds-averaged Navier-Stokes
DEM Discrete element method
DPM Discrete particle model

1. Introduction
In recent years, there has been an increase in the risk of wildfires in the forested areas of southwestern China [1–4]. However, the
complex terrain and low road density make it difficult for large ground-based firefighting equipment such as fire trucks to access the
fire sites. Aerial firefighting methods are often expensive and have limitations. Therefore, portable firefighting equipment is preferred
by forest fire departments in the region [5–7]. Among them, the pneumatic fire extinguisher is popular due to its simple operation,
economic applicability, and high efficiency compared to other portable firefighting tools.
In China, the 6 MF-30 pneumatic fire extinguisher is a commonly used forest and grassland firefighting equipment that generates
high-speed air flow through a two-stroke gas turbine to extinguish fire [8–10]. However, despite the effectiveness of the tool, forest
fires in China have been unpredictable, and the situation remains complicated and severe. The development of the 6 MF-30 pneumatic
fire extinguisher has stagnated in the past decade, and it is not efficient in fighting wildfire caused by thick-tree branches [7]. The
structure of the pneumatic fire extinguisher is simple, consisting of a two-stroke gasoline engine, a centrifugal impeller, and a casing, as
shown in Fig. 1(a). The engine provides power to the centrifugal impeller, which forces the air to flow at high speed. The casing

2
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

includes a volute and an air duct. The gas thrown out by the centrifugal impeller is collected by the volute and guided into the air duct.
The air duct has a rectangular cross-section at the end that connects it to the volute and a smaller circular end at the outlet. The
reduction in the size of the cross-sectional area enhances the initial velocity of the air jet.
Previous research has been done on sand throwing firefighting [11,12], which provide valuable references. Guo et al. combine the
sand-throwing equipment with EDEM full-size simulation to investigate its effectiveness in longitudinally extinguishing small-size
wood crib fires. Various tools, including HD cameras, infrared thermal imagers, CAD software, and armored K-type thermocouples,
were employed to record the fire extinguishment process, observe flame height changes, and monitor temperature variations. The
analysis is based on the energy conservation principle of the control volume, shedding light on the mechanism of the sand-throwing
equipment for extinguishing wood crib fires [11]. Yao et al. focus on the utilization of sand as a fire extinguishing material in ex­
periments involving diesel, ethanol, and n-heptane fuels. Through extensive experiments using a small-scale sand-throwing equip­
ment, the throwing performance of the equipment was tested and evaluated. The experiments included sand throwing to extinguish
circular pool fires, measuring and analyzing indicators such as flame shape, flame temperature, fire extinguishing time, and fire
extinguishing mechanism. The analysis revealed that covering the fuel to isolate combustibles and oxygen is considered the primary
fire extinguishing mechanism, with sand demonstrating excellent coverage and suffocation capabilities while minimizing the possi­
bility of re-ignition. The findings contribute to understanding the experimental phenomenon and mechanism of sand throwing to
extinguish pool fires, providing insights for optimizing the fire extinguishing performance of the equipment [12]. Some other literature
also provided valuable references for this study [13–17].
This study aims to improve the firefighting efficiency of the pneumatic fire extinguisher by enhancing it with a sand injection unit at
the air duct outlet to create pneumatic sandblasting fire extinguisher, as shown in Fig. 1(b). However, this optimization method has not
been proposed before, and there is no existing literature to refer to. Therefore, there is a need for a preliminary determination of the
external parameters of the extinguisher.
This study aims to investigate the sensitivity of geometric parameters of pneumatic sandblasting fire extinguishers to blasting
effects by conducting systematic research using experiments and CFD-DEM coupling simulation. Previous research work [18] analyzed
and simulated nine impeller designs through experiments and CFD-DEM coupling simulation. It found that the airflow field has a
significant impact on the sand spraying process in a wind-free environment and the analysis of sand ejection characteristics. Another
study [19] conducted a sensitivity analysis of sand ejection extinguisher parameters using DOE and CFD-DEM coupling simulation
technologies. The study used various methods like orthogonal experimental design, range and variance analysis, and single-factor
method to analyze the factors that affect the scattering distance and tilt degree of the equipment. Literature [20] optimized the
impeller of sand ejection fire extinguisher based on CFD-DEM simulation and Kriging model. Hence, their research provides a theo­
retical and methodological basis for the optimization design of this study. However, this study does not rely on direct collision of blades
to scatter sand, but rather it depends on the high-speed airflow in the air duct to carry and blast. Therefore, this study should pay more
attention to the drag force generated by the airflow. Considering the limited research on flow field simulations of this type of
pneumatic fire extinguisher, this study will refer to similar research in centrifugal fans. Hariharan and Govardhan [21], Madhwesh
et al. [22], and Karanth and Sharma [23] have studied fluid flow characteristics in centrifugal fans using the CFD technology. Younsi

Fig. 1. (a) 6 MF-30 pneumatic fire extinguisher, (b)Pneumatic sandblasting fire extinguisher.

3
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

et al. [24], through their research, analyzed the aerodynamics in the forward-swept centrifugal fan using the SST k-ω turbulent model.
In contrast, Petit and Nilsson found that the standard k-ε turbulence model was a more accurate predictor of results than the Realizable
k-ε, RNG k-ε, and SST k-ω turbulence models [25]. Chougule et al. performed CFD calculations of multiple air jet impacts on a flat plate
and found that the SST k-ω model can better predict the fluid flow and heat transfer characteristics [26–29]. This study will employ
multiple flow field calculations using these turbulence models and select the best one based on experimental results.
The main objective of this study is to use experiments and CFD-DEM coupled simulations to investigate the sensitivity of geometric
parameters of pneumatic sandblasting fire extinguishers to blasting effects. The findings of this study will have significant practical
implications for optimizing pneumatic fire extinguishers and enhancing fire-fighting efficiency in forest fire prevention and control. By
identifying the most sensitive geometric parameters and their optimal values, fire extinguisher manufacturers and fire-fighting
agencies can make informed decisions regarding design improvements. This knowledge can lead to the development of more effi­
cient and effective fire-fighting equipment, ultimately saving lives, protecting natural resources, and minimizing the devastating
consequences of forest fires.

2. Methods
2.1. Sandblasting experimental settings
A total of 12 sand collection boxes were positioned in neat, consecutive order at a distance of 1.5 m from the air duct outlet. Each
sand collection box housed 35 small collection boxes, each of which measured 6 cm × 6 cm × 4 cm. The angle of the extinguisher was
adjusted to a 20◦ downward tilt, and the speed was stabilized at approximately 4500 rpm by using a variable frequency controller.
In the experimental setup, after the airflow in the pneumatic sandblasting fire extinguisher was properly established and running
smoothly, a precisely measured amount of 3 kg of sand was used for the sandblasting process. This weight was selected to ensure a
sufficient quantity of sand for the experiment while maintaining consistency and repeatability. To inject the sand into the collection
boxes, a controlled release mechanism was employed. The sand was carefully poured or released into the sandblasting apparatus,
which directed the sand particles towards the target area. The collection boxes were strategically placed to capture the sand particles
blasted by the airflow. These collection boxes were designed to effectively collect and retain the sand, allowing for subsequent
measurements and analysis. After the sandblasting process was completed, the collection boxes were carefully collected and weighed
individually. The weights of the sand in each collection box were measured using precision weighing scales. This step enabled the
determination of the quality distribution of the sand that was successfully blasted to the ground. It is important to note that the ex­
periments were conducted with a high degree of control and precision. The sandblasting process was carefully calibrated, ensuring
consistent parameters and conditions throughout the experiments. The injection of the sand, the measurement of sand weights, and the
collection of data were performed with meticulous attention to detail to minimize any potential sources of error. To ensure the
reliability of the experimental results, appropriate measures were taken to maintain consistency and repeatability. Each experiment
was conducted multiple times, and the entire procedure was replicated to obtain an average and verify the consistency of the findings.
This approach allowed for the assessment of the reproducibility of the results and the identification of any potential variations or
outliers.
The experimental scene, as illustrated in Fig. 2, visually depicts the setup used for the experiments. It showcases the arrangement of
the sandblasting apparatus, the collection boxes, and the overall experimental environment, providing a clear representation of the
experimental setup and the context in which the sandblasting and sand collection occurred. By following these controlled procedures
and conducting repeatable experiments, the study ensures the accuracy and reliability of the data obtained, allowing for meaningful
analysis and conclusions regarding the sand distribution and the performance of the pneumatic sandblasting fire extinguisher.

2.2. Discrete element methods


In this study, the Discrete Element Method (DEM) is chosen as the computational method to analyze the motion and mechanical
behavior of materials. The Discrete Element Method (DEM) is a computational method that uses collision models to analyze material
behavior under various conditions. DEM was first proposed by British scholar P. Cundall based on the theory of spherical contact
mechanics, which has been widely used in various fields, including pharmaceuticals, mining, and agriculture [30].
In this paper, the collision model analyzing the motion and mechanical behavior of materials is the Hertz-Mindlin non-slip model,
which utilizes the Hertzian contact theory for the normal force component [31] and the Mindlin-Deresiewicz model for the tangential
force model [32]. Both forces also have damping components, with the damping coefficient associated with the recovery coefficient
[33]. The friction force follows Coulomb’s friction law [34], and the rolling friction force is achieved using the contact-independent

Fig. 2. Sandblasting experimental illustration.

4
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

directional constant torque model [35].


When two spherical particles 1 and 2 collide, the normal overlap εn is:

εn = r1 + r2 − |r→1 − →
r2 |, (1)

in which r1 and r2 represent the radius of two spherical particles, while → r represents the vector position of the particle center.
When two particles come into contact, the contact is in the form of a circular arc, with a contact radius of a, which can be calculated
using the following formula:
√̅̅̅̅̅̅̅̅
a = εn r∗ , (2)

in which r∗ represents equivalent radius of particles, which can be calculated by


1 1
r∗ = + . (3)
r1 r2
The normal force between particles, Fn , can be expressed as follows:
4
(4)
1 3
Fn = E∗ (r∗ )2 ε2n ,
3

in which E∗ represents equivalent modulus of elasticity, which can be expressed as:

1 1 − μ21 1 − μ22
∗ = + . (5)
E E1 E1

E1 , E2 and μ1 , μ2 the modulus of elasticity and Poisson’s ratio of two particles, respectively.
The formula for calculating the normal damping force Fnd between particles is as follows:
√̅̅̅
5 √̅̅̅̅̅̅̅̅̅̅ (6)
Fnd = − 2 β Sn m∗ vrn ,
6

where m∗ is the equivalent mass of the particle, vrn represents the normal component of relative velocity of the particle, β is the co­
efficient and Sn is the normal stiffness. The calculation formulas of m∗ , vrn , β and Sn are as follows:
1 1
m∗ = + , (7)
m1 m2

vrn = (v1 − v2 ) • →
n, (8)


r1 − →
r2

n = → →, (9)
|r1 − r2 |

ln e
β= , (10)
(ln e)2 + π 2

Sn = 2E∗ a, (11)

in which →n is the normal unit vector between two particles, and e is the coefficient of restitution.
Based on the above formula, the total normal contact force between particles can be obtained:
√̅̅̅̅̅̅̅̅̅̅̅̅̅
4 ∗ ∗ 12 3 2 ln e 5 ∗ ∗
E m (εn r∗ )4 (v1 − v2 ) • → (12)
1
Fn,12 = E (r ) εn −
2 n .
3 (ln e)2 + π2 3
The tangential force between particle 1 and particle 2 can be calculated using the following formula:
Ft = − St εt , (13)

in which, εt and St represent tangential overlap and stiffness, respectively. The value of St can be obtained from the following equation:
St = 8G∗ a, (14)

in which, G∗ is the equivalent shear modulus:

1 − μ21 1 − μ22
G∗ = + . (15)
G1 G2

5
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

G1 , G2 are the shear modulus of two particles.


The calculation formula for tangential damping force Ftd is as follows:
√̅̅̅
5 √̅̅̅̅̅̅̅̅̅
Ftd = − 2 β St m∗ vrt , (16)
6

in which vrt is the relative tangential velocity between two particles.


The formula for the total tangential contact force between particles is obtained from the above formula:
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
√̅̅̅̅̅̅̅̅ 2 ln e 20 ∗ ∗
(17)
1
Ft,12 = − 8G∗ r∗ εn εt − G m (εn r∗ )4 vrt .
(ln e)2 + π2 3
The tangential moment Tt can be calculated according to the following formula:
Tt = μs Fn (18)

where μs represents the static friction coefficient


In the simulation, rolling friction is also a key parameter. The formula for calculating the friction torque Tr on the contact surface is
as follows:

Tr = − μr Fn ri →
ωi (19)

where μr , ri , →
ωi represent the rolling friction coefficient, the distance between the center of mass and the contact point, and the unit
angular velocity vector of the particle at the contact point, respectively.

2.3. Drag force theories


In this study, drag force theories are employed to analyze the motion and resistance of particles in the fluid medium. These theories
play a crucial role in understanding and predicting the behavior of particles in different flow regimes.
According to the force acting on particles, the equation can be obtained as:
→ → →( )
du u − →up g ρp − ρ →
(20)
p
mp = mp + mp + F,
dt τr ρp

in which, mp represents the mass of particles, → u represents the velocity of the fluid, →
up is the velocity of particles, ρ is the density of the
→ →u− →up
fluid, ρp represents the density of particles, F is the additional force, and mp τr represents the drag force. τr represents the relaxation
time of particles, which can be calculated by:

ρp dp2 24
τr = , (21)
18μ Cd Re

where μ represents the viscosity of the fluid, dp represents the diameter of particles, and Re represents the relative Reynolds number.
Particle rotation is a natural part of particle motion and can have a significant impact on the trajectory of particles in the fluid. The
impact is more apparent for large and heavy particles with a high moment of inertia. In this case, if the particle rotation is ignored in
the simulation study, the generated particle trajectory may differ significantly from the actual particle path. To explain particle
rotation, it is necessary to solve the additional ordinary differential equation (ODE) for particle angular momentum:
( )
ωp ρf dp 5 ⃒⃒→⃒⃒ → →
d ̅→
Ip = Cω ⃒ Ω ⃒ • Ω = T, (22)
dt 2 2

in which, Ip represents the moment of inertia, ̅→


ωp represents the angular velocity of particles, ρf represents the density of the fluid, Cω
→ →
represents the rotational drag coefficient of particles, T represents the torque applied to the particles in the fluid domain, and Ω
represents the relative particle-fluid angular velocity, which can be expressed as:
→ 1
uf − →
Ω = ∇×→ ω p. (23)
2
For spherical particles, the moment of inertia Ip is calculated as follows:
π
Ip = ρp dp5 . (24)
60

when modeling and calculating in a moving reference frame, the force acting on particles around the axis defined for rotation can be
written in the Cartesian direction as:

6
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

( ) ( )
ρ ρ
mp 1 − Ω2 x + 2mp Ω up,y − uy (25)
ρp ρp
( ) ( )
ρ ρ
mp 1 − Ω2 y + 2mp Ω up,x − ux (26)
ρp ρp

where up,y and uy are the particle and fluid velocities in the y Cartesian direction, separately. up,x and ux are the particle and fluid
velocities in the x Cartesian direction, respectively.

2.4. Coupling simulation settings


CFD-DEM coupling theory, also known as the Coupled Computational Fluid Dynamics and Discrete Element Method, is a powerful
numerical simulation technique that combines the strengths of both methods to study complex multiphase flows. This approach en­
ables the accurate prediction of particle-fluid interactions and their impact on the overall flow behavior. In CFD-DEM coupling, the
CFD method is used to model the fluid phase, while DEM is employed to simulate the behavior of individual particles. By coupling these
two methods, researchers can gain a comprehensive understanding of the complex interplay between the fluid and particles in a wide
range of applications. One of the key advantages of CFD-DEM coupling is its ability to capture both the macroscopic and microscopic
phenomena simultaneously. The CFD component provides insights into the global flow characteristics, such as velocity profiles and
pressure distributions, while the DEM component offers detailed information about particle motion, collisions, and contact forces.
This coupling technique finds extensive applications in various fields, including pharmaceuticals, chemical engineering,
geotechnical engineering, and granular materials. It has proven particularly valuable in studying processes involving particle-fluid
interactions, such as fluidized beds, particle-laden flows, and sediment transport [18].
The effect of sand injection parameters and number of blades on the outlet airflow speed, as well as the distance of sand blasting,
was analyzed using Fluent and EDEM software. The simulation model was based on the Hertz no-slip model, various turbulence
models, and the DPM model. The geometric and mesh design, after independent validations, are presented in Fig. 3, respectively. The
results show that increasing the number of blades and modifying the sand injection parameters can significantly improve the blasting
effect. The application of the Hertz no-slip model and fluid dynamics principles in the simulation model has helped in understanding
the behavior of sand ejection.
In this study, Fluent is used to simulate the movement of high-speed airflow generated by the rotation of impellers. The

Fig. 3. (a) Geometry of the pneumatic fire extinguisher, (b) Mesh of internal flow field and external flow field of the extinguisher.

7
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

computational domain consists of four regions: the impeller region, volute region, air duct region, and the external flow field region.
The impeller calculation domain utilizes the multiple reference frame model (MRF), and the rotational speed is set to 4500 rpm. The
SIMPLE algorithm is employed in the solution strategy with pressure and momentum selected as Second Order, while turbulent kinetic
energy and turbulent dissipation rate as First Order Up Wind.
EDEM is used to simulate the sand particles generation and collisions during rotation in the air duct and external flow field. To
mimic experimental conditions closely, it is necessary to set the particle and equipment properties appropriately, as detailed in Table 1.
Sand particles are injected into the unit through the sand inlet with an initial falling speed of 1 m/s, while the injection flow rate is
controlled at 0.3 kg/s based on experimental data.
The two software exchange data for a coupled computation during calculations. The Discrete Phase Model (DPM) model in Fluent
uses the Euler-Lagrange method, treating the fluid phase as a continuous medium and allowing the discrete phase to exchange mass,
momentum, and energy with it. The software predicts the trajectory of discrete-phase particles through an integration method based
on particle force balance. In this study, data exchange was achieved by importing EDEM’s coupling interface to Fluent in the form of
UDF. Refer to Ref. [18] for specific coupling steps.

3. Validations
3.1. Validation of the flow field simulations
In order to determine the accuracy of the pneumatic fire extinguisher model establishment, grids generated, and solution strategy,
validation researches were conducted in this study. The standard k-ε, RNG k-ε, Realizable k-ε, and SST k-ω turbulence models were
utilized for calculation, and the turbulence model that was most consistent with the experiment was ultimately selected for further
research. In Fig. 4(a), the simulated results were compared with the experiment, where the yellow area signifies the range airflow
speeds measured at different positions. It was discovered that when utilizing the standard k-ε model for calculation, the airflow speeds
at each position simulated fell within the range of measured airflow speeds, while the calculated results of RNG k-ε, Realizable k-ε, and
SST k-ω models all had considerable discrepancies in comparison with the experiment. It shows that the results calculated by the
standard k-ε turbulence model were in agreement with the experimental results.
In the coupling simulation, the mesh size, boundary conditions, turbulent model, solution strategy, and other selections were
consistent with the pneumatic fire extinguisher. In Fig. 4(b), the experimental and simulated results are compared. It can be seen from
the figure that the calculated airflow velocity is slightly higher than the experimental results with a maximum error of 12.96% (at 2.5
m); the minimum error is 2.87% (at the outlet). The curve of the simulated airflow velocity change with the position is consistent with
the trend of the measured curve, and the error is within a tolerable range. Therefore, this simulation has reference value.

3.2. Validation of CFD-DEM coupling simulations


The existing sand injection unit has a sandblasting outlet with a widening angle of 8◦ and a diameter of 9 cm, and the length of the
air duct is 69 cm (10 cm of which belongs to the sand injection unit). While maintaining the sandblasting outlet diameter and the
overall length of the air duct, three different sandblasting outlet shapes were designed: flat, expanded, and narrow outlets, as depicted
in Fig. 5(a). Out of the three, the expanded outlet was subjected to experimental testing. It was observed that the injected sand hardly
contacted the lower wall of the air duct and instead propagated in the direction of the outlet and eventually fell into the collection box
under the influence of gravity, drag force, contact forces, and air resistance in the atmosphere. The total collected sand mass in the
range of 1.5 m–2.34 m being 1.3248 kg out of the 3 kg sand mass thrown out, covering an area of 1.68 × 0.9 m. All collection boxes are
capable of recovering 77.5% of the total sand mass, leaving only sparsely distributed unrecoverable sand outside the collection box. It
is evident that the placement of the collection box is the main concentration area of sand, and the area where sand can be effectively
stacked is within the 1.5–2.34 m range, with recoverable sand occupying 56.95% of the total sand mass thrown out.
In order to observe the distribution of sand on the ground, a contour of the collected sand mass distribution from all collection boxes
was created and weighed. It was found that the sand is distributed non-uniformly, with a core area displaying the most mass accu­
mulation in an agglomerated state. The mass distribution of sand outside the core area gradually decreases, and the distribution is also
more scattered. When referring to “sand scattering distance” in the following studies without specific instructions, it refers to the
position of the sand accumulation core area.
The simulated dispersion distance of the sand varies from 1.58 to 2.31 m, with a calculated core area length of 0.73 m and a width
of 0.28 m. The core area has a deviation of 6.67% in width, and 13.1% in length, while the dispersion distance deviates from 1.28% to
5.33%. These results indicate that the simulation can be effectively used for analysis.
From Fig. 5(b), it can be observed that a change to the sandblasting outlet configuration does indeed affect the spread distance of

Table 1
EDEM material properties settings.

Particle material parameters value Equipment material parameters value

Poisson’s ratio 0.35 Poisson’s ratio 0.269


Density (kg/m3) 2650 Density (kg/m3) 7890
Shear modulus (Pa) 7.692 × 107 Shear modulus (Pa) 8.117 × 107
Restitution coefficient (particle-particle) 0.01 Restitution coefficient (particle-particle) 0.01
Sliding friction coefficient (particle-particle) 0.5 Sliding friction coefficient (equipment-particle) 0.5
Rolling friction coefficient (particle-particle) 0.01 Rolling friction coefficient (equipment-particle) 0.01

8
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

Fig. 4. (a) Airflow speed under different turbulence models, (b)Comparison of experimental and simulated airflow speed of pneumatic sandblasting flow field.

the sand. When the sandblasting outlet is changed to a flat configuration, the spread distance of the sandblasting is observed to be
between 1.76 and 2.6 m, with the core area length and width being 0.84 m and 0.36 m, respectively. This change represents an 11.39%
increase in the nearest end and a 12.56% increase in the farthest end compared to the expanded outlet. A narrow outlet configuration
results in a sandblasting distance of between 2.06 and 2.54 m, with a core area length of 0.48 m and a width of 0.18 m. Although this
configuration significantly increases the blasting distance, the increase in the coverage area is minimal. In conclusion, from the
perspective of sandblasting distance, the narrow outlet is the most effective configuration, followed by the flat outlet, and the expanded
outlet has the shortest distance. From the perspective of sand coverage, the flat outlet has the largest coverage area, followed by the
expanded outlet, and the narrow outlet has the smallest coverage area. The primary influence of the three different-shaped outlets on
sandblasting distance comes from the changes in airflow speed. As shown in Fig. 5, outlet airflow speed can significantly impact the
sandblasting distance as a related drag force for carrying sand movement. Although the narrow outlet has a higher outlet airflow speed,
the coverage area is smaller. The flat outlet has an appropriate outlet airflow speed and a wider coverage area. The subsequent
optimization can be focused on achieving the characteristics of high outlet airflow speed and wide sand coverage.

4. Results and discussion


4.1. Effects of the air duct length
The length of the air duct is an important factor in determining the velocity of the sand jet and its distribution. To investigate the
effect of air duct length on sand distribution, the original length of 69 cm was changed to 59 cm, 64 cm, 74 cm, and 79 cm. At a fixed
rotational speed, the distribution of sand and the distance of the sand jet were studied for each air duct length. The sand distribution
after blasting was found to be more concentrated for a 74 cm duct length, but the distribution was highly uneven. On the other hand,

9
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

Fig. 5. (a)Different shapes of the sand injection unit, (b) Impact of different mouths on the sand throwing distance, (c) Impact of different mouths on the flow field.

for 59 cm and 64 cm duct lengths, the sand was evenly distributed but with shorter coverage. When the length of the air duct was
increased to 79 cm, the sand was uniformly distributed over a greater distance, with the accumulation area being between 1.69 and
2.77 m away from the air duct.
The velocity distribution of the sand group during sandblasting was analyzed, and it was observed that the sand was more scattered
at the sampling position for 59 cm and 64 cm duct lengths, but with a larger velocity. Conversely, for 74 cm and 79 cm duct lengths, the
sand was more concentrated in the same position, although the velocity was smaller. This information is depicted in Fig. 6(a).
The size and distribution of the outlet jet velocity are directly affected by changes in the length of the air duct. Consequently, this
leads to differences in the distance and dispersion state of sand. Fig. 6(b) illustrates the outlet airflow speed distribution under different
air duct lengths. As the length of the air duct decreases, the high-speed area significantly increases, and the average airflow speed
improves. This conclusion is consistent with the phenomenon in Fig. 6(a). However, it’s important to note that increasing the outlet
airflow velocity is not always better. After extending the length of the air duct, the kinetic energy carried by the sand is no longer
sufficient. Consequently, most of the sand falls into the core area without secondary motion, which achieves better accumulation and
coverage.
Based on the above analysis, extending the length of the air duct to 79 cm carried sand to a farther area without causing violent
rebounds, making it a good trade-off despite the decrease in average airflow speed and kinetic energy. Additionally, most particles still
fall in the core area with uniform distribution, increasing the blasting distance by 12.67%–18.38% compared to the original pneumatic
sandblasting extinguisher. Therefore, the determined length of the air duct is 79 cm.

10
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

Fig. 6. (a) Velocity and distribution of sand particles in vertical slices under different lengths of the air duct, (b)Airflow velocity distribution at the outlet under
different air duct lengths.

4.2. Effects of the sandblasting diameter


The sandblasting diameter influences the distribution of sand. The sandblasting apparatus has four designed diameters, which are 6
cm, 8 cm, 10 cm, and 12 cm, while the existing sandblasting diameter is 9 cm. Of these, the sandblasting unit with diameters of 6 cm
and 8 cm is internal embedded in the original air duct, while the others are external embedded. The air velocity at the respective outlet
and at distances of 1 m, 1.5 m, 2 m, and 2.5 m from the outlet were measured by a Pitot tube. As shown in Fig. 7(a), the air velocity at
the outlet under the four diameters is not lower than 53 m/s. However, once the high-speed airflow jet enters the atmospheric
environment, the velocity attenuation slows at 1.5–2.5 m from the outlet after rapidly decreasing at a distance of 1–1.5 m. As the
diameter increases, the outlet airflow speed and the jet airflow speed along the flow path decrease sequentially.
Fig. 7(b) shows the distribution of sand under different sandblasting diameters. It can be observed that the sandblasting distance
decreases with an increase in diameter. Fig. 7(c) illustrates the curve of sand concentration on the central axis with a shift in distance.
Since an increase in sandblasting diameter results in a decrease in airflow velocity, the sandblasting distance of sand under the internal
embedded method is greater than that under the external method, and a significant difference exists. The sandblasting range under the
internal embedded method is between 1.75–2.85 m, while it is between 1.5–2.5 m under the external method. Note that the maximum
heap density of sand on the ground reaches 610 kg/m2 under the external embedded, indicating that a larger sandblasting diameter
enables the concentration of sand. From the perspective of sandblasting distance, when the diameter is small, the airflow velocity is
high, and the sand is blasted farther. Considering all factors, it is believed that the sandblasting effect of the internal embedded method
is better than that of the external embedded method, and the best sandblasting diameter is determined as 8 cm.

4.3. Effects of the blade number


Research has indicated that the distance of sand sandblasting is primarily determined by the airflow speed at the outlet. While
changing the parameters of sand blasting can indirectly affect the airflow volume, the effect is limited. The blade, which is the core
component of the fan, determines both the airflow speed and the range of distance. By optimizing the blade parameters, it is possible to
further increase both the airflow volume at the outlet of the fan as well as the range distance of sand sandblasting. This can ultimately
improve the effectiveness of the fire extinguisher.
To determine the optimal blade parameters, research was done on the pneumatic sandblasting fire extinguisher with a flat

11
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

Fig. 7. (a) Outlet and airflow jet speed under different sandblasting diameters, (b) Sand distribution under different sandblasting diameters, (c)Sand distribution in the
midline under different sandblasting diameters.

12
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

sandblasting outlet. Different blade numbers were set, including 14, 16, 18, 20, and 22, with a sandblasting diameter of 8 cm and an air
duct length of 79 cm. Fig. 8(a) displays the jet velocity under different blade numbers, revealing that the outlet airflow speed is at its
highest with 20 blades. Furthermore, the velocity retained along the way is also the highest with 20 blades.
A simulation was performed to observe the effect of blade numbers on sandblasting. The results are shown in Fig. 8(b), which show
that when there were 20 blades, the sand blasting distance range was at its highest, ranging from 1.89 to 3.03 m, achieving the farthest
distance. This observation indicates that the outlet airflow speed is of great importance in determining the range of sand sandblasting.
This implies that by increasing the airflow speed at the outlet, the range of sand sandblasting can be expanded. Nonetheless, having
excessively high speed can cause the sand group to travel too fast, making it difficult to cover the accumulation effectively. After
optimizing the parameters, the sandblasting distance increased by 19.62%–31.17%.

4.4. Fire extinguishing effects of the improved pneumatic sandblasting fire extinguisher
Wood stacks were used as solid fire sources in the fire extinguishing experiment. The wood stack was assembled from four layers of
pine strips with a cross section of 4cm × 4 cm, and two sizes were set up. The length of the pine strips of the small stack was 25 cm, and
that of the large stack was 30 cm. It was found in the experiment that when the wood stack was placed farthest at a distance of 2.5 m
from the air duct, it exceeded the effective sandblasting range of 1.5–2.34 m and the effective airflow jet distance of 1.5 m of the
original pneumatic sandblasting fire extinguisher. Furthermore, the airflow speed at this location was lower than 15 m/s, which
belongs to the range of combustion-supporting airflow speed, thus resulting in fire extinguishing failure.
After adjusting various parameters, the optimized pneumatic sandblasting extinguisher increased its effective airflow jet distance to
1.9 m and its effective sandblasting range to 1.89–3.03 m. The airflow speed at a distance of 2.5 m exceeded 15 m/s. In order to verify
that the optimized extinguisher could effectively increase the extinguishing distance and extinguishing efficiency, fire extinguishing

Fig. 8. (a) Velocity of airflow along the jet path with different blade numbers, (b) Sandblasting distance under different blade numbers.

13
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

experiments were conducted again after optimization. The small-sized and large-sized wood stacks were placed at a distance of 2.5 m
from the air duct, and the optimized extinguisher was used for extinguishing after the wooden stacks were freely ignited for 90 s. It was
found that the optimized extinguisher could effectively extinguish both the small-sized and large-sized wood stacks, as shown in Fig. 9
(a) and (b). The extinguishing time of the small wooden stack was 238 s and that of the large wooden stack was 267 s.
The average airflow speed measured at a distance of 3 m from the air duct outlet is around 8.1 m/s, which is too low to extinguish
fires. Additionally, it may even contribute to the fire. A small-sized wood stack was placed at a distance of 3 m from the outlet to test the
firefighting effect relying on the sandblasting. As shown in Fig. 9(c), after 386 s, it was discovered that the sand could still successfully
extinguish the stack, but it took a longer time. In the end, it required a quality of 14.31 kg for the sand to extinguish the fire, fully
demonstrating the effectiveness of the optimization.

5. Conclusions
This study identified optimal design parameters for pneumatic sandblasting fire extinguishers through a combination of simula­
tions and experiments, which yielded the following conclusions:
Firstly, narrow sandblasting outlets exhibit high airflow velocities but result in limited sand coverage areas. On the other hand, flat
outlets demonstrate moderate airflow velocities, the farthest sand ejection distance, and wider coverage areas. Expanded outlets, in
contrast, have lower airflow velocities and shorter sand ejection distances.
Secondly, an increase in the length of the air duct leads to a decrease in the average airflow velocity at the outlet. The dispersion of
sand particles is more significant when the air duct is short, accompanied by high sand velocity. As the air duct length increases, sand
particles become more concentrated within the airflow, resulting in reduced velocity. Notably, when the air duct reaches a length of 79
cm, sand particles are evenly piled up, resulting in the largest sand coverage area. Compared to the original pneumatic sandblasting
fire extinguisher, the sandblasting distance increases by 12.67%–18.38% under these optimized conditions.
Thirdly, an increase in the diameter of the sandblasting outlet promotes the concentration of sand particles, while a decrease in
diameter enhances the sandblasting distance. Through comprehensive comparison, it is observed that setting the sandblasting outlet
diameter to 8 cm yields a sandblasting distance ranging from 1.75 to 2.85 m, with a maximum accumulation of 610 kg/m2. The blade
count of the impeller also influences the sandblasting results, with the optimal blade count being 20. Under this optimal parameter
combination, the sandblasting distance increases by 19.62%–31.17%.
Lastly, the improved pneumatic sandblasting fire extinguisher demonstrates a longer fire extinguishing distance compared to the
original extinguisher. The original extinguisher was unable to extinguish the fire of a small-sized wood stack located 2.5 m away.
However, after optimization, the fire of the small-sized wood stack located at the same distance was effectively extinguished within
238 s, while it took 267 s for a large-sized wood stack. Additionally, the optimized extinguisher successfully extinguished the fire of a
small-sized wood stack located 3 m away from the air duct outlet, demonstrating the effectiveness of the proposed improvement
method.
It is important to mention that this findings of this study on optimal design parameters for pneumatic sandblasting fire extin­
guishers are subject to certain limitations. The identified conclusions are specific to the experimental setup and conditions utilized in
this study, and may not directly apply to other fire extinguisher designs or environments. Further research is necessary to explore the

Fig. 9. Fire extinguishing experiments: (a) Small-sized wood stack at 2.5 m, (b) Big-sized wood stack at 2.5 m, (c) Small-sized wood stack at 3.0 m

14
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

applicability of these results in various contexts, accounting for different fire scenarios and extinguisher configurations.

Authorship contributions
Category 1.
Conception and design of study: Fanbao Chen, Tingting Xu; acquisition of data: Fanbao Chen, Guanyu Hou, Tao Deng, Zhenhua
Jiang; analysis and/or interpretation of data: Fanbao Chen, Guanyu Hou, Tingting Xu, Ziyang Wang.
Category 2.
Drafting the manuscript: Fanbao Chen, Guanyu Hou, Guoqing Zhu; revising the manuscript critically for important intellectual
content: Fanbao Chen, Guanyu Hou,Tingting Xu, Jianhua Huang.
Category 3.
Approval of the version of the manuscript to be published (the names of all authors must be listed): Fanbao Chen, Tingting Xu,
Guanyu Hou, Jianhua Huang, Guoqing Zhu, Tao Deng, Zhenhua Jiang, Ziyang Wang.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

Data will be made available on request.

Acknowledgements
This study is funded by the China Scholarship Council (202206420073), Key Technology Research and Development Program of
Shandong (CN) [2021CXGC011303], the Central Universities under Grant (2022QN1010 and 2022QN1011), and the Natural Science
Foundation of Jiangsu Province (BK20221124).

References
[1] F.J. Diao, L.F. Shu, X.R. Tian, M.Y. Wang, Change trends of forest fire danger in Yunnan Province in 1957-2007, Chinese Journal of Ecology 28 (11) (2009)
2333.
[2] X. Tian, F. Zhao, L. Shu, M. Wang, Hotspots from satellite monitoring and forest fire weather index analysis for southwest China, Forest Research, Beijing 23 (4)
(2010) 523–529.
[3] X. Tian, L. Shu, F. Zhao, M. Wang, Forest fire danger changes for southwest China under future scenarios, Sci. Silvae Sin. 48 (1) (2012) 121–125.
[4] X.R. Tian, F.J. Zhao, L.F. Shu, M.Y. Wang, Changes in forest fire danger for south-western China in the 21st century, Int. J. Wildland Fire 23 (2) (2014) 185–195.
[5] Xuemei Yao, Analysis on current situation and trend of forest fire prevention technology at home and abroad, Agriculture and Technology 36 (6) (2016), 173+
213. (in Chinese).
[6] Wei Maozhou, Keyin Wang, Current situation and prospect of forest fire fighting equipment, For. Mach. Woodwork. Equip. (2006), 2006 (07): 11-14. (in
Chinese).
[7] Shuanglei Zhu, Research on Aerodynamic Performance of Axial Flow Impeller of Axial Flow Pneumatic Extinguisher, Beijing Forestry University, Beijing, 2010.
[8] S.Q. Zhang, G.S. Yu, Research and development of portable pneumatic extinguishers, For. Mach. Woodwork. Equip. 40 (6) (2012) 8–10.
[9] L. Wang, Y. Sun, W. Yu, H. Yu, Y. Yang, Portable pneumatic extinguisher automatic testing platform, Journal of Forestry Engineering 1 (1) (2016) 105–110.
[10] L. Wang, W. Li, J. Chen, Ergonomic evaluation of the operating characteristics of the 6MF-30 portable pneumatic extinguisher, Appl. Ergon. 51 (2015) 39–43.
[11] W. Guo, G. Zhu, B. Yao, F. Chen, X. Xu, Study on the fire extinguishing mechanism of small size wood crib based on small sand-throwing equipment, Case Stud.
Therm. Eng. 25 (2021), 100942.
[12] B. Yao, G. Zhu, W. Guo, F. Chen, Z. Wang, X. Xu, Experimental study of the effectiveness of sands on extinguishing pool fires based on small sand-throwing
equipment, Case Stud. Therm. Eng. 28 (2021), 101408.
[13] A.C. Huang, C.F. Huang, Z.X. Xing, J.C. Jiang, C.M. Shu, Thermal hazard assessment of the thermal stability of acne cosmeceutical therapy using advanced
calorimetry technology, Process Saf. Environ. Protect. 131 (2019) 197–204.
[14] A.C. Huang, C.F. Huang, Y. Tang, Z.X. Xing, J.C. Jiang, Evaluation of multiple reactions in dilute benzoyl peroxide concentrations with additives using
calorimetric technology, J. Loss Prev. Process. Ind. 69 (2021), 104373.
[15] A.C. Huang, Z.P. Li, Y.C. Liu, Y. Tang, C.F. Huang, C.M. Shu, J.C. Jiang, Essential hazard and process safety assessment of para-toluene sulfonic acid through
calorimetry and advanced thermokinetics, J. Loss Prev. Process. Ind. 72 (2021), 104558.
[16] H.L. Zhou, J.C. Jiang, A.C. Huang, Y. Tang, Y. Zhang, C.F. Huang, C.M. Shu, Calorimetric evaluation of thermal stability and runaway hazard based on
thermokinetic parameters of O, O–dimethyl phosphoramidothioate, J. Loss Prev. Process. Ind. 75 (2022), 104697.
[17] M.L. Zhang, X.L. Dong, Y. Tang, A.C. Huang, F. Chen, Q.C. Kang, Z.X. Xing, Experimental investigations of extinguishing sodium pool fires using modified
expandable graphite powders, Case Stud. Therm. Eng. 32 (2022), 101911.
[18] F. Chen, G. Zhu, C. Wang, W. Shang, B. Yao, W. Guo, X. Xu, Sand-ejecting fire extinguisher parameter sensitivity analysis based on experiments and CFD-DEM
coupling simulations, Powder Technol. 395 (2022) 443–454.
[19] F. Chen, G. Zhu, B. Yao, W. Guo, T. Xu, Sand-ejecting fire extinguisher parameter sensitivity analysis based on DOE and CFD-DEM coupling simulations, Adv.
Powder Technol. 33 (9) (2022), 103719.
[20] F. Chen, G. Zhu, X. Wang, B. Yao, W. Guo, T. Xu, D. Cheng, Optimization of the impeller of sand-ejecting fire extinguisher based on CFD-DEM simulations and
Kriging model, Adv. Powder Technol. 34 (1) (2023), 103898.
[21] M. Heinrich, R. Schwarze, Genetic optimization of the volute of a centrifugal compressor, in: 12th European Conference on Turbomachinery Fluid Dynamics &
Thermodynamics, EUROPEAN TURBOMACHINERY SOCIETY, 2017.
[22] N. Madhwesh, K. Vasudeva Karanth, N. Yagnesh Sharma, Investigations into the flow behavior in a nonparallel shrouded diffuser of a centrifugal fan for
augmented performance, J. Fluid Eng. 140 (8) (2018).
[23] K.V. Karanth, N.Y. Sharma, CFD analysis on the effect of radial gap on impeller-diffuser flow interaction as well as on the flow characteristics of a centrifugal
fan, Journal of Rotating Machinery (2009), 2009.

15
F. Chen et al. Case Studies in Thermal Engineering 49 (2023) 103354

[24] M. Younsi, F. Bakir, S. Kouidri, R. Rey, Influence of impeller geometry on the unsteady flow in a centrifugal fan: numerical and experimental analyses, Int. J.
Rotating Mach. (2007), 2007.
[25] O. Petit, H. Nilsson, Numerical investigations of unsteady flow in a centrifugal pump with a vaned diffuser, Int. J. Rotating Mach. (2013), 2013.
[26] N.K. Chougule, G.V. Parishwad, P.R. Gore, S. Pagnis, S.N. Sapali, CFD analysis of multi-jet air impingement on flat plate, in: Proceedings of the World Congress
on Engineering, vol. 3, 2011, July, 2078-0958.
[27] S.S. Thakre, J.B. Joshi, CFD modeling of heat transfer in turbulent pipe flows, AIChE J. 46 (9) (2000) 1798–1812.
[28] Y. Shi, M.B. Ray, A.S. Mujumdar, Computational study of impingement heat transfer under a turbulent slot jet, Ind. Eng. Chem. Res. 41 (18) (2002) 4643–4651.
[29] T.H. Shih, W.W. Liou, A. Shabbir, Z. Yang, J. Zhu, A new k-ε eddy viscosity model for high Reynolds number turbulent flows, Comput. Fluids 24 (3) (1995)
227–238, https://doi.org/10.1016/0045-7930(94)00032-T.
[30] P.A. Cundall, O.D. Strack, A discrete numerical model for granular assemblies, Geotechnique 29 (1) (1979) 47–65.
[31] H. Hertz, The contact of elastic solids, J Reine Angew, Math 92 (1881) 156–171, https://doi.org/10.1515/crll.1882.92.156.
[32] R.D. Mindlin, H.E. Deresiewicz, Elastic Spheres in Contact under Varying Oblique Forces, 1953.
[33] Y. Tsuji, T. Tanaka, T. Ishida, Lagrangian numerical simulation of plug flow of cohesionless particles in a horizontal pipe, Powder Technol. 71 (3) (1992)
239–250, https://doi.org/10.1016/0032-5910(92)88030-L.
[34] P.A. Cundall, O.D. Strack, A discrete numerical model for granular assemblies, Geotechnique 29 (1) (1979) 47–65, https://doi.org/10.1680/geot.1979.29.1.47.
[35] H. Sakaguchi, E. Ozaki, T. Igarashi, Plugging of the flow of granular materials during the discharge from a silo, Int. J. Mod. Phys. B 7 (09n10) (1993)
1949–1963, https://doi.org/10.1142/S0217979293002705.

16

You might also like